Next Article in Journal
Kinetic and Thermodynamic Analysis of High-Pressure CO2 Capture Using Ethylenediamine: Experimental Study and Modeling
Next Article in Special Issue
Laminar Burning Velocities of Hydrogen-Blended Methane–Air and Natural Gas–Air Mixtures, Calculated from the Early Stage of p(t) Records in a Spherical Vessel
Previous Article in Journal
Decarbonizing Local Mobility and Greenhouse Agriculture through Residential Building Energy Upgrades: A Case Study for Québec
Previous Article in Special Issue
Reactivity Model as a Tool to Compare the Combustion Process in Aviation Turbine Engines Powered by Synthetic Fuels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of the Multiple Injection Strategy on Combustion and Emission Characteristics of a Two-Stroke Marine Engine

1
South Sea Fisheries Management Service, 128 Imhang-ro, Jeju-si 63280, Korea
2
Graduate School of Korea Maritime and Ocean University, 727, Taejong-ro, Yeongdo-gu, Busan 49112, Korea
3
Division of Marine System Engineering, Korea Maritime and Ocean University, 727, Taejong-ro, Yeongdo-gu, Busan 49112, Korea
4
Interdisciplinary Major of Maritime AI Convergence, Korea Maritime and Ocean University, 727, Taejong-ro, Yeongdo-gu, Busan 49112, Korea
*
Authors to whom correspondence should be addressed.
Energies 2021, 14(20), 6821; https://doi.org/10.3390/en14206821
Submission received: 17 September 2021 / Revised: 11 October 2021 / Accepted: 14 October 2021 / Published: 19 October 2021
(This article belongs to the Special Issue Numerical Investigations of Combustion)

Abstract

:
This paper presents research on the effects of the multiple injection strategies on the combustion and emission characteristics of a two-stroke heavy-duty marine engine at full load. The ANSYS FLUENT simulation software was used to conduct three-dimensional simulations of the combustion process and emission formations inside the engine cylinder in both single- and double-injection modes to analyze the in-cylinder pressure, temperature, and emission characteristics. The simulation results were then compared and showed good agreement with the measured values reported in the engine’s sea-trial technical reports. The simulation results showed reductions in the in-cylinder pressure and temperature peaks by 6.42% and 12.76%, while NO and soot emissions were reduced up to 24.16% and 68%, respectively, in the double-injection mode in comparison with the single-injection mode. However, the double-injection strategy increased the CO2 emission (7.58%) and ISFOC (23.55%) compared to the single-injection. These are negative effects of the double-injection strategy on the engine that the operators need to take into consideration. The results were in line with the literature reviews and would be good material for operators who want to reduce the engine exhaust gas emission in order to meet the stricter IMO emission regulations.

Graphical Abstract

1. Introduction

In comparison with gasoline engines, direct injection (DI) diesel engines offer better reliability and higher thermal efficiency. Therefore, they have been being widely and commonly used not only in light-duty public transport, private vehicles but also in heavy-duty marine transportation. However, because of the increasing environmental hazards of marine engine exhaust emissions, as well as the International Maritime Organization (IMO) emission regulations being more and more stringent, the exhaust gas emissions such as nitrogen oxide (NOx), soot, and carbon dioxide (CO2) must be reduced to meet the stricter emission standards [1].
Theoretically, engine exhaust gas emission reduction approaches are divided into three categories, including (1) pre-treatment; (2) internal treatment; and (3) post-treatment [2]. In the pre-treatment approach, fuels, the main factor that affect the engine’s combustion and emission characteristics, are substituted or handled prior to supply to engines. Regarding substitute fuels, various gas fuels are currently being increasingly employed in marine engines to reduce engine exhaust gas emissions. Effects of various alternative gas fuels on reducing engine exhaust gas emissions were analyzed in detail in our previous publications [3,4].
In the internal treatment approach, the combustion of the fuel inside the engine cylinder will be modified towards the goal of achieving lower exhaust gas emissions. Using innovative fuel injection systems [5,6], lean or rich combustions [7], pre-chamber type combustion, innovative combustion bowl feature [8], etc. are examples of this method. In addition, the method of adding water to the engine combustion chamber to reduce the peak temperature in the cylinder or modifying the engine working cycle (exhaust gas recirculation–EGR for example) are also other effective ways to reduce emissions. They are also being effectively used in the marine industry today.
In the post-treatment approach, it is necessary to equip exhaust gas after-treatment devices to treat the exhaust gas after it has left the engine. Using selective catalytic reduction (SCR) systems for NOx reduction is an example of this method [2]. These methods don’t require modifications on fuels or existing engines, but exhaust gas after-treatment de-vices are mandatory. This results in an increase in additional costs and makes the engine room become more cramped.
Focusing on the internal treatment approach, multiple injection strategies have been proven to be effective solutions to reduce soot or even NOx and soot simultaneously in both heavy-duty DI diesel engines [9,10,11,12,13] and light-duty DI diesel engines [14,15,16,17,18,19,20,21,22,23]. Nehmer and Reitz [9] showed that multiple injection helps to reduce both particle matter (PM) and NOx emissions from controlled pressure rise. The effectiveness of multiple injection strategies for the simultaneous reduction of NOx and PM in DI diesel engines was also reported in [10,11,13,14,15,16,18,19,20,21,22]. Numerical simulations were also carried out to explore the chemical mechanism of NOx and soot reduction for multiple injection strategies by Han et al. [11]. They found that, in multiple-injection strategies, due to the subsequent injections occurred in a high-temperature environment that was produced by combustion of previous injections, the injected fuel burnt more rapidly resulting in a reduction in soot formation rates. Therefore, the net production of soot was dramatically reduced. Furthermore, soot formed in a high-temperature environment had been oxidized more effectively leading to a great reduction in engine-out soot emissions in multiple injection strategies. These previous researches have proven that multiple injection strategies have similar effects to the retarded injection strategy in the single injection method on NOx reduction applied on conventional DI diesel engines. The reduction in soot emissions is due to the fact that the soot-producing fuel-rich regions are significantly reduced when the fuel dose is divided into several times to be injected into the combustion chamber. Zhang [14] reported that the soot emission was closely related to the pilot flame (first-dose injection flame) and could be reduced by reducing this pilot flame by using only a small amount of fuel for the first injection in multiple injection strategies. Another study on multiple injection strategies was done by Tanaka et al. [16]. They also reported that it is possible to simultaneously reduce exhaust gas emissions by decreasing the influence of the pilot flame through minimizing the fuel quantity in the pilot (first-time) injection or retarding this first-time injection timing.
Meanwhile, Choi and Reitz [17] experimentally studied the effects of single and split injections on soot formation in a DI engine. They found that the split injection reduced soot emissions without penalty on NOx emissions compared with the single injection. Montgomery and Reitz [12] also reported the same finding when they evaluated the effects of EGR and multiple injections on PM and NOx emissions from a DI diesel engine. Yang and Chung also reported in [23] that, when multiple injection strategies were applied, a significant reduction in soot emissions could be achieved without penalty on NOx emissions.
Based on the literature above, it is clear that the multiple injection strategy offers an effective solution to reduce engine exhaust gas emissions. This study focuses on the combustion process and emission characteristics inside the cylinder of a two-stroke marine main engine equipped on a university’s training ship.
The simulation software ANSYS Fluent with its state-of-the-art models was used to perform 3D simulations calculating the combustion process and emission formation inside the engine cylinder. There are two operating modes of the engine that have been simulated, including the single injection mode and the double injection mode, in order to analyze the engine in-cylinder pressure, temperature, and emission characteristics. The simulation results were then validated by comparing them with the engine’s sea-trial data.

2. Simulated Engine and CFD Analysis

2.1. Simulated Engine Specifications

The research object of this study is a 2-stroke diesel marine engine. The combustion process of the engine is studied and simulated to investigate the effects of the double-injection strategy on the engine’s in-cylinder pressure, temperature, and emission characteristics compared to those when using the conventional single-injection strategy. Two engine operating modes were simulated: (1) single-injection mode and (2) double-injection mode. In both the two modes, apart from changes in the injection strategy, other operating parameters, and thus simulation parameters were kept unchanged in order to clearly investigate the effects of the injection strategy on the engine output and emissions.
The three-dimensional (3D) model of the engine cylinder, piston, cylinder head with exhaust valve of the simulated engine is shown in Figure 1. The engine has six cylinders with only one exhaust valve on each cylinder mounted at the cylinder head center and two fuel injectors mounted symmetrically on the cylinder head. The piston surface is a U-type shape. The engine cylinder diameter and piston stroke are 420 mm and 1360 mm, respectively. Specifications of the simulated engine are shown in Table 1.

2.2. CFD Analysis

The ANSYS Fluent 2019R2 simulation software with its state-of-the-art models was employed to simulate the combustion and emission formation inside the engine cylinder. The simulation process was performed from the exhaust valve closing (EVC) to the exhaust valve opening (EVO) and included three steps: (1) pre-processing, (2) processing, and (3) post-processing. The first step included building the computational domain (Geometry), creating movable computational meshes (Meshing), and setting up the simulation parameters (Setup). The solutions were calculated in the second step. In the third step, the simulation results were analyzed and reported. After obtaining the simulation results, the simulation results and measured results were compared for the validation of the simulation models. This process was repeated until the simulation and measured results matched. Figure 2 shows the workflow of the study.
The k- ε turbulence model was selected to simulate the turbulence of the fluid flow inside the engine cylinder. This is the commonly widely used model in the simulation of the combustion of internal combustion engines (ICEs) [24]. To simulate the combustion process of the engine the Non-premixed Combustion model was used. This model allows simulating non-premixed combustion in which fuels and oxidizers are supplied to the combustion chamber separately [25]. The Discrete Phase Model (DPM) was used to simulate the fuel injection. This model offers a way to correct the injection velocities and initial droplet diameters due to cavitation. In order to simulate the breaking up of the phenomenon of fuel droplets, the Kelvin-Helmholtz Rayleigh-Taylor (KHRT) sub-model was used. The Auto-ignition model was used for simulating the self-ignition phenomenon that occurs in an auto-ignited diesel engine.
The Extended Zeldovich model was used to simulation the NO formation characteristic [25,26]. This chemical reaction mechanism consists of seven species and three reactions and is able to predict NO formation with high accuracy over a wide range of equivalence ratios. This model was presented in detail and can be found in Appendix C of our previous publication [3,4]. The Moss-Brookes model [25,27] was used to simulate the soot formation due to the combustion process. To simulate the collisions and interactions between fuel droplets and walls the Stochastic Collision sub-model, which is based on the algorithm of O’Rourke [25], was used. The non-premixed combustion model and Moss-Brookes soot models can be found in references [25,27]. Other CFD models can be found in references [25,26,27]. Table 2 shows a summary of the numerical models used in this study.

2.3. Computational Mesh, Boundary and Initial Conditions

The combustion chamber geometry was built by using the AutoCAD software (Autodesk, California, USA) and the computational mesh was then created by the ANSYS (Autodesk, California, USA) Meshing platform. The boundary conditions (BCs), initial conditions, and simulation parameters were declared in the setup platform. The solution was then calculated in series using an 8-core processor and took approximately 72 h of CPU time. Figure 3 shows half of the 3D symmetry combustion chamber geometry while Figure 4 shows half of the 3D computational mesh of the combustion chamber at the top dead center (TDC).
Due to the movement of the piston, we needed to use a dynamic mesh in this study. Due to the re-meshing during calculation, the mesh quality and mesh resolution might be changed. In order to keep the mesh quality within acceptable range, we used the dynamic layering method with a collapse factor of 0.4 and a split factor of 0.4 for dynamic moving mesh modeling. The computational mesh will be re-meshed according to the movement of the piston based on the dynamic layering method. Specifically, during the piston downward movement, mesh layers will be added when the outermost layer (boundary piston surface mesh layer) height increases by more than 40% (split factor of 0.4) of the declared height when setting the meshing. In contrast, during the piston upward movement, mesh layers will be collapsed when the outermost layer height decreases by more than 40% (collapse factor of 0.4) of the declared height. This means that the number of mesh layers and thus the number of elements will be increased or decreased according to the moving down or up of the piston. In this way, the mesh quality and resolution will be kept almost unchanged during calculation.
Due to the specificity of the dynamic layering method, we have used hexahedral elements for computational mesh. This type of elements is also the most perfect for computational mesh due to its very high orthogonal quality. We had also performed a lot of simulations with different mesh resolutions (coarse, medium, and fine) until mesh-independence of the calculation results was achieved. Finally, we obtained the final mesh with parameters as shown in Table 3.
The BCs at the cylinder head, piston surface and cylinder liner were defined as impermeable wall BCs. The cylinder geometry was planed symmetric, so the cyclic BCs were assigned to the cutting surfaces of the computational domain. The simulation was started from the EVC of 30 CAD after the bottom dead center (ABDC) (EVC = 30 CAD ABDC) to the EVO of 30 CAD before the bottom dead center (BBDC) (EVO = 30 CAD BBDC). The schematic of the engine working cycle is shown in Figure 5 while Figure 6 shows the start of injection (SOI) and injection duration in both the single and double injection modes.
The BCs and initial conditions were chosen from the sea-trial report of the actual engine provided by the operating officer. Table 4 shows the BCs and initial conditions for the numerical simulation cases.

2.4. Simulation Cases and Fuel Properties

In this study, the combustion process of the engine running in two fuel injection modes at full load was simulated, including single injection and double injection. In single injection mode, fuel was injected directly into the engine’s combustion chamber in just one injection. Meanwhile, the fuel was divided into two injections to inject into the engine’s combustion chamber in double injection mode. All other simulation conditions were kept unchanged. The properties of the diesel fuel are shown in Table 5.
The simulation results were then validated against the engine’s sea-trial data provided by the engine manufacturer. The simulation results showed a good agreement with the measured results written in the engine’s sea-trial technical reports and were presented in detail in Section 3–Simulation results.

3. Simulation Results

To clearly investigate the effects of the injection strategy on the combustion and emission characteristics of the engine, all of the BCs and engine’s working conditions were kept unchanged, excepting the difference in the fuel injection strategy. In this section, the in-cylinder pressure, temperature and emissions characteristics of the engine will be presented and analyzed.

3.1. In-Cylinder Pressure

The simulated in-cylinder pressure diagrams in both injection modes are shown in Figure 7a while Figure 7b shows the comparison between simulated and experimental peak pressure inside the engine cylinder obtained from the sea-trial. As can be seen in Figure 7, the peak pressures in both single and double injections are similar to those in comparison with the experimental results. The deviation between simulated and experimental results in the single injection mode was only 4.85%, while it was 8.3% in the double injection mode. These deviations are acceptable in CFD analysis. Comparing the two simulation cases, the results showed that the in-cylinder peak pressure in the double injection mode was 6.42% lower than that compared with the single injection mode (Figure 7a).
As shown in Figure 6, in the double injection mode, since only 25% of the total fuel was injected into the engine cylinder before the TDC and 75% of the total fuel was injected from the TDC, the rate of heat release (RoHR) in the engine cylinder before the TDC in the double injection mode was lower than that in the single-injection mode. From the TDC, even though the fuel injected into the engine cylinder in the double injection mode was higher than that in the single-injection mode, but the fuel was injected after the TDC when the piston was in the downward movement. These resulted in a reduction in the in-cylinder peak pressure in the double injection mode compared with the single-injection mode, as shown in Figure 7a. The reductions in the in-cylinder peak pressure when using multiple injection strategy were also reported by Nehmer and Reitz in [9], and Qiu et al. in [22].

3.2. In-Cylinder Temperature

The in-cylinder temperature diagrams with RoHR traces in both simulation cases are shown in Figure 8. The simulation result showed that the in-cylinder peak temperature in the double-injection mode was lower than that in comparison with the single-injection mode. However, the mean in-cylinder temperature in the double-injection mode was higher than that compared with the single-injection mode during the expansion process. This resulted in a higher exhaust gas temperature in the double-injection mode compared to the single-injection mode.
As the same as in-cylinder peak pressure, the decrease in the in-cylinder peak temperature in double-injection compared with single-injection is the result of less fuel being injected before the TDC, resulting in lower RoHR inside the engine cylinder. On the other hand, more fuel was injected into the engine cylinder after the TDC in the double-injection mode resulting in a higher RoHR inside the engine cylinder and thus higher in-cylinder temperature during the expansion process as can be seen in Figure 8.

3.3. NO Emission

The NO emissions generated in both injection modes are shown in Figure 9a, while Figure 9b shows the comparison between simulated and experimental NO emission of the engine obtained from the sea-trial. As can be seen in Figure 9a, the NO emission generated in the double injection mode was 24.16% lower than that compared with the single injection mode. Figure 9b shows a very good agreement between Simulated and experimental NO emission in both injection modes. The deviation between simulated and experimental results in the single injection mode was only 3.6%, while it also was only 3.62% in the double injection mode.
According to the extended Zelodovich NO formation mechanism, the NO formation is greatly influenced by the in-cylinder temperature peak and oxygen (O2) concentration within the engine cylinder. NO formation occurs in regions in the cylinder where the local temperature is above 1800 K and the formation rate increase significantly with the increase of the local in-cylinder temperature [30,31,32]. The simulation results showed a reduction of 12.76% in the in-cylinder temperature peak in the double-injection mode compared to the single-injection mode (Figure 8). This resulted in a reduction of 24.16% in NO emission in the double-injection mode in comparison with the single-injection mode as shown in Figure 9a. Reduction effects on NOx emission as well as soot (PM) have been reported by Fang et al. in their publication on Fuel, 2008 [18] and other researchers in [9,10,11,13,14,15,16,19,20,21,22].

3.4. Soot Formation

Soot is the main component of particulate matter (PM) emission [33,34,35,36]. Under high temperature and equivalence ratios (fuel-rich conditions), which are typically found in diesel diffusion combustions, hydrocarbon fuels have a strong tendency to form carbonaceous particles, i.e., soot. Soot particles are formed very early in the diffusion combustion process because of the dissociation of fuels under high temperature and high equivalence ratio conditions. Usually, under normal engine running conditions of diesel engines, most of the soot that formed in the early combustion stage will be depleted due to oxidation in the late stages of the combustion. This typically takes place in oxygen-rich regions of the combustion chamber in the late stage of the engine cycle (expansion process). Therefore, in diesel engines, the soot formation, and completeness of the soot oxidation process determines the particle emission characteristics of the engine [34]. The parameters that play the most important role during the soot formation inside the engine cylinder are the local equivalence fuel/air ratio (C/H-ratio and C/O-ratio), in-cylinder temperature, pressure, and residence time [34].
Under ideal reaction conditions, the products of hydrocarbon fuel combustions are only CO2 and H2O. The amount of oxygen required for a complete chemical reaction (the stoichiometric oxygen requirement- O 2 ,   s t o i c h i o m e t r i c ) can be calculated from the following reaction equation [25]:
C n H m O k + ( n + m 4 k 2 ) O 2 n CO 2 + m 2 H 2 O
In the above equation k, m, and n represent the number of oxygen, hydrogen and carbon atoms of the considered fuel, respectively.
The actual amount of oxygen available for combustion is expressed by the air access ratio λ or by its inversion, the equivalence (fuel-air) ratio ϕ :
ϕ = 1 λ = O 2 ,   s t o i c h i o m e t r i c O 2 ,   a c t u a l
The soot formations in both injection modes are shown in Figure 10. The simulation results showed that the maximum soot emission generated in the double-injection mode was reduced up to more than 68%. In addition, the soot mass fraction plot in the double-injection mode showed two peaks as can be seen in Figure 10. This is the result of the double-injection strategy in which the fuel dose was divided into two injections to be injected into the engine cylinder.
Reduction in soot formation is explained as the reduction effect in the fuel/air equivalence ratio ϕ. By dividing the fuel dose into two injections, the mass fraction of fuel in each injection was decreased resulting in a reduction in the fuel/air equivalence ratio, the factor that mainly and strongly affects the soot formation in ICEs as widely known. Figure 10 also showed that, in the double-injection mode, the amount of soot produced inside the engine cylinder was not only reduced but they were also completely oxidized more quickly in comparison with the single-injection mode. This is because the oxidation of soot in the cylinders has been promoted by higher cylinder temperatures during the engine expansion process in the double-injection. Once again, reduction effects on soot as well as NOx emission when using double-injection strategy is in line with the report by Fang et al. in their publication on Fuel, 2008 [18] and other researchers in [9,10,11,12,13,14,15,16,17,19,20,21,22,23].
In addition, Tree et al. mentioned in [37] that fuel pyrolysis plays an important role in soot formation in compression-ignition engines (CIEs). Pyrolysis is a process that changes the molecular structure of organic compounds, such as fuels, under high-temperature conditions without significant oxidation although O2 species may be present. All fuels will undergo pyrolysis and produce basically the same species: polyacetylenes, polycyclic aromatic hydrocarbons (PAH), unsaturated hydrocarbons, and especially acetylene. Pyrolysis reactions are usually endothermic owing to the fact that their rates are strongly temperature-dependent [38]. Fuel pyrolysis leads to the production of some species which are precursors or building blocks for soot. Fuel pyrolysis to form soot precursor increase with temperature. Therefore, soot increases as the temperature increases. The same tendency was also reported in [39,40]. In this study, as presented in Figure 8, the in-cylinder peak temperature was reduced in the double-injection strategy. As a result, this resulted in a decrease in soot formation inside the engine cylinder, as shown in Figure 10.

3.5. Carbon Dioxide (CO2) Emission

The CO2 mass fraction in both injection modes is shown in Figure 11a, while Figure 11b shows the comparison between simulated and experimental CO2 emission of the engine obtained from the sea-trial. The simulation results pointed that the CO2 emission in the case of double-injection was 7.58% higher than that of single-injection (Figure 11a). In comparison with the experimental results, the deviation between the simulated and experimental CO2 mass fraction was 7.36% (Figure 11b) and acceptable in CFD analysis.
Regarding the reasons for the increasing CO2 emission in the double-injection mode, as has been known, carbon dioxide is a product of the complete combustion of hydrocarbon fuels. Firstly, hydrocarbon fuels will be oxidized to CO during the combustion process, and then it is oxidized to form CO2 sequentially if the in-cylinder temperature is high enough and there is still enough O2 in the cylinder. Therefore, CO2 formation strongly depends on the in-cylinder temperature and the concentration of oxygen in the engine cylinder. As can be seen in Figure 8, even though the in-cylinder peak temperature in the double-injection mode was lower than that in the single-injection mode, but the in-cylinder temperature during expansion stroke was higher because most of the fuel (75% of mass) was injected and burned after the TDC, in the late combustion phase of the double-injection mode. The higher temperature inside the engine cylinder during the late stage of the combustion in the case of double-injection mode promoted CO oxidation to produce CO2 resulting in an increase in CO2 emission in the double-injection mode in comparison with the single-injection mode, as shown in Figure 11a.

3.6. Specific Fuel Oil Consumption

Figure 12 shows the indicated specific fuel oil consumption (ISFOC) in both injection modes. The simulation results showed an increase in the ISFOC of the engine when operated in the double-injection mode. Specifically, the ISFOC was increased 23.55% when operating the engine in the double-injection mode compared to that in the single-injection mode. The increase in the ISFOC can be explained by a decrease in the peak pressure in the cylinder leading to a decrease in engine power when operated in the double-injection mode. This result points that using the double-injection strategy to reduce NO and soot emission not only results in an increase in CO2 emission but also an increase in the SFOC of the engine. These are negative effects of the double-injection strategy on the engine that the operators need to take into consideration when operating the engine.

4. Conclusions

This paper numerically investigated the effects of the double-injection strategy on the combustion and emission characteristics of a two-stroke diesel marine engine. The simulation results were then compared and showed good agreement with the measured results obtained from the engine’s sea-trials.
The main outcomes of the study are as follows:
  • The in-cylinder peak pressure in the double-injection mode was 6.42% lower compared with the single injection mode.
  • The in-cylinder peak temperature in the double-injection mode was lower in comparison with the single-injection mode resulting in a significant reduction in NO emission (24.16%).
  • The maximum soot emission generated in the double-injection mode was reduced by more than 68% in comparison with the single-injection mode.
  • However, the double-injection strategy increased the CO2 emission (7.58%) and ISFOC (23.55%) compared to the single-injection. These are negative effects of the double-injection strategy on the engine that the operators need to take into consideration when operating the engine.
The results were in line with the existing literature and provides good material for operators who want to reduce engine exhaust gas emissions in order to meet the stricter IMO emission regulations.

Author Contributions

Conceptualization, V.C.P. and W.-J.L.; Methodology, V.C.P. and W.-J.L.; Software, V.C.P.; Validation, J.-H.S. and W.-J.L.; Formal analysis, J.-H.S.; Data curation, J.-H.S. and W.-J.L.; Writing—original draft preparation, V.C.P.; Writing—review and editing, V.C.P. and J.-H.S.; Project administration, W.-J.L.; Funding acquisition, W.-J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Ministry of Education of Republic of Korea and the National Research Foundation of Korea (NRF-2019R1G1A1005342) and by the Ministry of Science and ICT of Republic of Korea and National IT Industry Promotion Agency (S1226-21-1001).

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.

References

  1. Thomson, H.; Corbett, J.J.; Winebrake, J.J. Natural gas as a marine fuel. Energy Policy. 2015, 87, 153–167. [Google Scholar] [CrossRef] [Green Version]
  2. Woodyard, D. Pounders Marine Diesel Engines and Gas Turbines, 9th ed.; Elsevier: Burlington, MA, USA, 2009. [Google Scholar]
  3. Pham, V.C.; Choi, J.H.; Rho, B.S.; Kim, J.S.; Park, K.N.; Park, S.K.; Le, V.V.; Lee, W.J. A numerical study on the combustion process and emission characteristics of a natural gas-diesel dual-fuel marine engine at full load. Energies 2021, 14, 1342. [Google Scholar] [CrossRef]
  4. Pham, V.C.; Rho, B.-S.; Kim, J.-S.; Lee, W.-J.; Choi, J.-H. Effects of various fuels on combustion and emission characteristics of a four-stroke dual-fuel marine engine. J. Mar. Sci. Eng. 2021, 9, 1072. [Google Scholar] [CrossRef]
  5. Martin, G.C.; Mueller, C.J.; Milam, D.M.; Radovanovic, M.S.; Gehrke, C.R. Early direct-injection, low-temperature combustion of diesel fuel in an optical engine utilizing a 15-hole, dual-row, narrow-included-angle nozzle. SAE Int. J. Engines 2008, 1, 1057–1082. [Google Scholar] [CrossRef]
  6. Polonowski, C.J.; Mueller, C.J.; Gehrke, C.R.; Bazyn, T.; Martin, C.M.; Lillo, M. An experimental investigation of low-soot and soot-free combustion strategies in a heavy-duty, single-cylinder, direct-injection, optical diesel engine. SAE Int. J. Fuels Lubr. 2012, 5, 51–77. [Google Scholar] [CrossRef]
  7. Dunn-Rankin, D. Lean Combustion Technology and Control; Elsevier Ltd.: Amsterdam, The Netherlands; Academic Press: Boston, MA, USA, 2011. [Google Scholar]
  8. Belgiorno, G.; Boscolo, A.; Dileo, G.; Numidi, F.; Pesce, F.C.; Vassallo, A.; Ianniello, R.; Beatrice, C.; Blasio, G.D. Experimental study of additive-manufacturing-enabled innovative diesel combustion bowl features for achieving ultra-low emissions and high efficiency. SAE Int. J. Adv. Curr. Pract. Mobil.-V130-99EJ 2021, 3, 672–684. [Google Scholar] [CrossRef]
  9. Nehmer, D.A.; Reitz, R.D. Measurement of the effect of injection rate and split injections on diesel engine soot and NOx emissions. SAE Pap. 1994, 940668. [Google Scholar]
  10. Tow, T.C.; Pierpont, D.A.; Reitz, R.D. Reducing particulate and NOx emissions by using multiple injections in a heady duty DI diesel engine. SAE Pap. 1994, 940897. [Google Scholar]
  11. Han, Z.; Uludogan, A.; Hampson, G.J.; Reitz, R.D. Mechanism of soot and NOx emission reduction using multiple injection in a diesel engine. SAE Pap. 1996, 960633. [Google Scholar]
  12. Montgomery, D.T.; Reitz, R.D. Six-mode cycle evaluation of the effect of EGR and multiple injections on particulate and NOx emissions from a DI diesel engine. SAE Tech. Pap. 1996, 960316. [Google Scholar]
  13. Pierpont, D.A.; Montgomery, D.T.; Reitz, R.D. Reducing particulate and NOx using multiple injections and EGR in a D.I. diesel. J. Fuels Lubr. 1995, 104, 171–183. [Google Scholar]
  14. Zhang, L. A study of pilot injection in a DI diesel engine. SAE Pap. 1999, 1, 3493. [Google Scholar]
  15. Chen, S.K. Simultaneous reduction of NOx and particulate emissions by using multiple injections in a small diesel engine. SAE Pap. 2000, 1, 3084. [Google Scholar]
  16. Tanaka, T.; Ando, A.; Ishizaka, K. Study on pilot injection of DI diesel engine using common rail injection system. JSAE Rev. 2002, 23, 297–302. [Google Scholar] [CrossRef]
  17. Choi, C.Y.; Reitz, R.D. An experimental study on the effects of oxygenated fuel blends and multiple injection strategies on DI diesel engine emissions. Fuel 1999, 78, 1303–1317. [Google Scholar] [CrossRef]
  18. Fang, T.; Coverdill, R.E.; Lee, C.F.; White, R.A. Effects of injection angles on combustion processes using multiple injection strategies in an HSDI diesel engine. Fuel 2008, 87, 3232–3239. [Google Scholar] [CrossRef]
  19. Chacko, N.; Rajkumar, S.; Thanragaja, J. Experimental and modeling analysis of multiple-injection strategies with B20 operation in a CRDI engine. Fuel 2021, 293, 120433. [Google Scholar] [CrossRef]
  20. Plamondon, E.; Seers, P. Parametric study of pilot–main injection strategies on the performance of a light-duty diesel engine fueled with diesel or a WCO biodiesel–diesel blend. Fuel 2019, 236, 1273–1281. [Google Scholar] [CrossRef]
  21. Jorques Moreno, C.; Stenlaas, O.; Tunestal, P. Influence of small pilot on main injection in a heavy-duty diesel engine. SAE Tech. Pap. 2017, 1, 708. [Google Scholar] [CrossRef]
  22. Qiu, L.; Cheng, X.; Liu, B.; Dong, S.; Bao, S.Z. Partially premixed combustion based on different injection strategies in a light-duty diesel engine. Energy 2016, 96, 155–165. [Google Scholar] [CrossRef]
  23. Yang, S.Y.; Chung, S.H. An experimental study on the effects of high-pressure and multiple injection strategies on DI diesel engine emissions. In SAE Technical Paper Series; SAE International: Warrendale, PA, USA, 2013. [Google Scholar]
  24. Durbin, P.A. Near-wall turbulence closure modeling without “damping functions”. Theor. Comput. Fluid Dyn. 1991, 3, 1–13. [Google Scholar]
  25. ANSYS Fluent Theory Guide; ANSYS, Inc.: Canonburg, PA, USA, 2013; Available online: http://www.ansys.com (accessed on 7 July 2021).
  26. Rao, V.; Honnery, D. A comparison of two NOx prediction schemes for use in diesel engine thermodynamic modelling. Fuel 2013, 107, 662–670. [Google Scholar] [CrossRef]
  27. Brookes, S.; Moss, J. Predictions of soot and thermal radiation properties in confined turbulent jet diffusion flames. Combust. Flame. 1999, 116, 486–503. [Google Scholar] [CrossRef]
  28. PubChem Compound. n-heptane–Compound Summary. National Center for Biotechnology Information. 8600 Rockville Pike, Bethesda MD, 20894 USA. 16 September 2004. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/8900 (accessed on 7 September 2020).
  29. Friend, D.G.; Ely, J.F. and Ingham, H. Thermophysical properties of methane. J. Phys. Chem. Ref. Data 1989, 18, 583. [Google Scholar] [CrossRef]
  30. Heywood, J.B. Internal Combustion Engine Fundamentals; McGraw-Hill Education: New York, NY, USA, 1998. [Google Scholar]
  31. Wei, L.; Geng, P. A review on natural gas/diesel dual fuel combustion, emissions and performance. Fuel Process. Technol. 2016, 142, 264–278. [Google Scholar] [CrossRef]
  32. Kuo, K.K. Principles of Combustion, 2nd ed.; John Wiley and Sons: New York, NY, USA, 2005. [Google Scholar]
  33. Maricq, M.M.; Chase, R.E.; Xu, N.; Laing, P.M. The effects of the catalytic converter and fuel sulfur level on motor vehicle particulate matter emissions: Light duty diesel vehicles. Environ. Sci. Technol. 2002, 36, 283–289. [Google Scholar] [CrossRef]
  34. Burnett, R.T.; Cakmak, S.; Brook, J.R.; Krewski, D. The role of particulate size and chemistry in the association between summertime ambient air pollution and hospitalization for cardiorespiratory diseases. Environ. Health Perspect. 1997, 105, 614–620. [Google Scholar] [CrossRef]
  35. Kittelson, D.B.; Watts, W.F.; Johnson, J.P. On-road and laboratory evaluation of combustion aerosols -part 1: Summary of diesel engine results. J. Aerosol Sci. 2006, 37, 913–930. [Google Scholar] [CrossRef]
  36. Rounce, P.; Tsolakis, A.; York, S.A.P.E. Speciation of particulate matter and hydrocarbon emissions from biodiesel combustion and its reduction by after-treatment. Fuel 2012, 96, 90–99. [Google Scholar] [CrossRef]
  37. Tree, D.R.; Svensson, K.I. Soot processes in compression ignition engines. Prog. Energy Combust. Sci. 2007, 33, 272–309. [Google Scholar] [CrossRef]
  38. Smith, O.I. Fundamentals of soot formation in flames with application to diesel engine particulate emissions. Prog. Energy Combust. Sci. 1981, 7, 275–291. [Google Scholar] [CrossRef]
  39. Gowthaman, S.; Sathiyagnanam, A.P. Effects of charge temperature and fuel injection pressure on HCCI engine. Alex. Eng. J. 2016, 55, 119–125. [Google Scholar] [CrossRef] [Green Version]
  40. Pan, W.; Yao, C.; Han, G.; Wei, H.; Wang, Q. The impact of intake air temperature on performance and exhaust emissions of a diesel methanol dual fuel engine. Fuel 2015, 162, 101–110. [Google Scholar] [CrossRef]
Figure 1. Structure principle schematic of the simulated engine.
Figure 1. Structure principle schematic of the simulated engine.
Energies 14 06821 g001
Figure 2. The workflow of the study.
Figure 2. The workflow of the study.
Energies 14 06821 g002
Figure 3. Half of the 3D symmetry combustion chamber geometry.
Figure 3. Half of the 3D symmetry combustion chamber geometry.
Energies 14 06821 g003
Figure 4. Half of the 3D computational mesh of the combustion chamber at the TDC.
Figure 4. Half of the 3D computational mesh of the combustion chamber at the TDC.
Energies 14 06821 g004
Figure 5. The schematic of the engine working cycle (TDC: Top dead center; BDC: Bottom dead center; EVO: Exhaust valve opens; EVC: Exhaust valve closes; SPO: Scavenging-air port opens; SPC: Scavenging-air port closes; SOI: Start of injection; EOI: End of injection).
Figure 5. The schematic of the engine working cycle (TDC: Top dead center; BDC: Bottom dead center; EVO: Exhaust valve opens; EVC: Exhaust valve closes; SPO: Scavenging-air port opens; SPC: Scavenging-air port closes; SOI: Start of injection; EOI: End of injection).
Energies 14 06821 g005
Figure 6. Injection properties in the operating modes.
Figure 6. Injection properties in the operating modes.
Energies 14 06821 g006
Figure 7. Simulated cylinder pressure diagrams (a) and comparison between simulated and experimental peak pressure (b).
Figure 7. Simulated cylinder pressure diagrams (a) and comparison between simulated and experimental peak pressure (b).
Energies 14 06821 g007
Figure 8. Simulated cylinder temperature diagrams with RoHR traces in both simulation cases.
Figure 8. Simulated cylinder temperature diagrams with RoHR traces in both simulation cases.
Energies 14 06821 g008
Figure 9. The mean NO mass fraction diagrams (a) and comparison between simulated and experimental NO emission (b).
Figure 9. The mean NO mass fraction diagrams (a) and comparison between simulated and experimental NO emission (b).
Energies 14 06821 g009
Figure 10. The mean soot mass fraction diagrams.
Figure 10. The mean soot mass fraction diagrams.
Energies 14 06821 g010
Figure 11. The mean CO2 emission diagrams (a) and comparison between simulated and experimental CO2 emission (b).
Figure 11. The mean CO2 emission diagrams (a) and comparison between simulated and experimental CO2 emission (b).
Energies 14 06821 g011
Figure 12. The ISFOC of the engine in both injection modes.
Figure 12. The ISFOC of the engine in both injection modes.
Energies 14 06821 g012
Table 1. The specifications of the simulated engine.
Table 1. The specifications of the simulated engine.
ParameterValueUnit
Type of Engine2-Stroke Diesel Engine
Name of Engine6L42MC-ME
No. of Cylinder6
Cylinder Bore × Stroke420 × 1360mm
Connecting Rod Length1700mm
Compression Ratio13.5
Rated Power @ Rated Speed4485 @ 160 kW @ rpm
IMEP18bar
No. of Fuel Injector2
FuelDiesel (C10H22)
LCV of Diesel42,343 *kJ/kg
* Provided by ANSYS Fluent database.
Table 2. The summary of the numerical models.
Table 2. The summary of the numerical models.
ModelDescription
CombustionNon-premixed Combustion model [25]
EmissionNOExtended Zeldovich
SootMoss-Brookes model
AtomizationSpray Discrete Phase Model (DPM)
BreakupKHRT
IgnitionAuto-ignition
CollisionStochastic Collision model
Turbulencek- ε turbulence model
Table 3. Mesh Statistics and Quality.
Table 3. Mesh Statistics and Quality.
Mesh Statistics and Quality
Type of elementsHexahedral
No. of nodes at the TDC459,058
No. of elements at the TDC508,650
Aspect ratio17.05
Maximum Skewness ratio0.76
Minimum Orthogonal quality0.45
Table 4. Boundary conditions and initial conditions for the numerically simulation cases.
Table 4. Boundary conditions and initial conditions for the numerically simulation cases.
Boundary ConditionsBoundary Type/Specific Condition
Cylinder headWall/Temp./297 °C
PistonMesh movement/Temp./297 °C K
LinerWall/Temp./197 °C
Segment cutCyclic
Initial conditionsValue
Scavenging air Press.2.0 bar
Scavenging air Temp.28 °C
EVC30 CAD ABDC
EVO30 CAD BBDC
SOI12 CAD BTDC
Injection duration32 CAD
Table 5. The diesel fuel properties [25,28,29].
Table 5. The diesel fuel properties [25,28,29].
FuelDieselUnit
Density730kg/m3 @ 25 °C, 1 at
Chemical name/formulaDiesel/C10H22
Lower calorific value42,000kJ/kg
Molecular weight142.284kg/kmol
Boiling point174°C
Auto-Ignition temperature223°C
Specific heat capacity (Cp)2090J/kg·K @ 25 °C
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Seol, J.-H.; Pham, V.C.; Lee, W.-J. Effects of the Multiple Injection Strategy on Combustion and Emission Characteristics of a Two-Stroke Marine Engine. Energies 2021, 14, 6821. https://doi.org/10.3390/en14206821

AMA Style

Seol J-H, Pham VC, Lee W-J. Effects of the Multiple Injection Strategy on Combustion and Emission Characteristics of a Two-Stroke Marine Engine. Energies. 2021; 14(20):6821. https://doi.org/10.3390/en14206821

Chicago/Turabian Style

Seol, Ju-Hwan, Van Chien Pham, and Won-Ju Lee. 2021. "Effects of the Multiple Injection Strategy on Combustion and Emission Characteristics of a Two-Stroke Marine Engine" Energies 14, no. 20: 6821. https://doi.org/10.3390/en14206821

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