Vous êtes sur la page 1sur 7

1st MARINELIVE International Workshop on Propulsion Systems January 11-12, 2012, Athens, Greece

CFD MODELING AND OPTIMIZATION OF MARINE DIESEL ENGINES

Lambros Kaiktsis Department of Naval Architecture and Marine Engineering National Technical University of Athens Heroon Polytechniou 9, 157 73 Athens, Greece kaiktsis@naval.ntua.gr

Christos Chryssakis Det Norske Veritas, Reseach and Innovation Veritasveien 1, 1363 Hovik, Norway christos.chryssakis@dnv.com

ABSTRACT
Recent legislation imposes strict limits on the emissions from new marine Diesel engines, especially in the Emission Control Areas (ECAs). Computational Fluid Dynamics (CFD), supported by experimental data, is a cost effective approach for characterizing engine flow and combustion processes. CFD, properly coupled with optimization tools, can be a valuable tool for guiding development towards cleaner and more efficient marine engines, complying with the recent regulations. This paper summarizes recent CFD research activities at the National Technical University of Athens aiming at understanding and optimizing marine Diesel engine aerothermochemistry. The examples discussed include injection profile optimization, Heavy Fuel Oil modeling, and water addition studies.

KEYWORDS
CFD, marine Diesel engines, optimization, Heavy Fuel Oil modeling, water injection.

1. INTRODUCTION
Recent attempts to reduce atmospheric pollution from seagoing ships include the adoption of a corresponding strategy by the European Commission [1]. Further, the International Maritime Organization (IMO) adopted in 1997 a set of regulations, outlined in Annex VI of the MARPOL Convention. MARPOL Annex VI came into force in 2005 and was modified in 2008, setting limits on the emissions of sulphur oxides (SOx) and nitrogen oxides (NOx) [2]. MARPOL Annex VI sets limits on SOx and NOx emissions, and also contains provisions for setting up special SOx Emission Control Areas (ECAs), which are characterized by more stringent controls on emissions. Along these lines, the Marine Environment Protection Committee, at its 57th session in April 2008, agreed on a three-tier structure, which would set progressively tighter NOx emission standards for new marine engines, depending on the date of their installation (see Figure 1). The Tier III standards presented in Figure 1 will be enforced in the ECAs

only, which include the North Sea, the Baltic Sea, and near-urban areas. Soot emission standards have not been announced yet but are expected to come into force in the near future. Regarding large two-stroke marine Diesel engines, it is evident from Figure 1 that, in comparison to recent engine designs, 2011 engines correspond to reduced NOx emissions by 15 per cent, while designs providing an 80 per cent reduction should be developed by 2016. Recent advances in Computational Fluid Dynamics (CFD) enable an accurate description of flow and combustion in marine Diesel engines. Given the lower cost of computational studies to that of experimental ones, validated CFD simulations can nowadays be a valuable tool in guiding engine development. CFD simulations could be combined with optimization techniques, for optimizing engine performance and emissions. Hence, such studies have been recently initiated. The present paper summarizes recent CFD research performed at the National Technical

University of Athens (NTUA), supervised by the authors, aiming at understanding and optimizing flow and combustion processes in large marine Diesel engines. The research examples presented here include: (a) the optimization of injection profile in a large two-stroke marine Diesel engine, (b) the development of Heavy Fuel Oil (HFO) models for CFD studies in marine Diesel engines, and (c) water addition studies for nitric oxides (NOx) reduction.

fuel oil consumption (SFOC). Thus, proper objective functions have been introduced, in particular: (a) the final NOx concentration, normalized by the corresponding value of a reference design of continuous injection, and (b) an approximation of SFOC, given by the ratio of the total injected fuel mass to the work integral of the closed part of the engine cycle (this ratio is also normalized by the corresponding reference value). Three problem set-ups were considered: (i) an unconstrained problem, (ii) a problem constrained by the maximum acceptable cylinder pressure (150 bar), and (iii) a problem constrained by both the maximum pressure and the minimum work output per engine cycle (that of the reference design of continuous injection). Bore diameter (m) Stroke (m) Injection system Exhaust valve closing (oATDC) Engine speed (rpm) 0.58 2.416 Common Rail -96 105

Exhaust valve opening (oATDC) 120


Figure 1. NOx emissions standards (g/kW h) of marine engines versus engine speed (rpm), according to the IMO [2].

Table 1. Main characteristics of the engine used in the optimization study of [3].

2. RESEARCH EXAMPLES
2.1 INJECTION PROFILE OPTIMIZATION IN A LARGE TWO-STROKE ENGINE Modern marine Diesel engines are equipped with Common Rail injection systems, allowing for a wide variation of the injection profile parameters. Recently, an optimization study was carried out at NTUA, aiming at optimizing the injection profile of a large two-stroke marine Diesel engine at full load [3]. The study has built on coupling a KIVA-3 based CFD code [4] with the optimization code EASY [5], which is based on evolutionary algorithms. The goal of this research has been the identification of optimal injection profiles, consisting in a pilot and a main injection event, for minimizing both NOx emissions and engine specific fuel consumption. The main engine characteristics are summarized in Table 1. The injection profiles considered in [3] consisted of a pilot and a main injection event, as sketched in Figure 2. These profiles have been parametrized in terms of four designs variables (see Table 2). Tetradecane (C14H30) has been used as the fuel. The goal of optimization has been the simultaneous minimization of NOx emissions and engine specific
2

By evolving an initial population of randomly generated individuals (possible solutions, each corresponding to a certain combination of the problem design variables), the genetic algorithm converges to the final set of optimal solutions, forming the Patero front. In this approach, individuals are evaluated based on the results of a CFD run. The outcome of all three problems considered in [3] is presented in Figure 3, in terms of the corresponding Pareto fronts. Evidently, both in the unconstrained and the one-constraint cases, the optimum solutions are characterized by substantial improvements in both NOx emissions (of the order of 1520 per cent) and SFOC (of the order of 2 percent). The improvements are marginal when both constraints are imposed. While the obtained NOx reduction of the order of 1520 per cent is substantial and complies with the 2011 regulations, it is far from meeting the 2016 standards. Thus, a more complex combination of incylinder measures, possibly including water addition and exhaust gas recirculation (EGR) is necessary to meet the 2016 standards with incylinder only measures (thus avoiding exhaust gas aftertreatment).

Design variable Start Of Pilot Injection - SOPI (oCA ATDC) Start Of Main Injection SOMI (oCA ATDC) Pilot Mass Fraction PMF (% of total injected mass) Mass reduction MR (% of reference design)

Min. value -40

Max. value -5

2.2 HEAVY FUEL OIL MODELING Marine Diesel engines commonly operate with Heavy Fuel Oil (HFO). HFO is a mixture of residual and lighter distillation products. HFO composition may vary substantially, thus introducing uncertainties in CFD modeling. An accurate representation of HFO thermophysical properties is essential for modeling fuel spray breakup and evaporation. The dynamics of both the primary (jet) and secondary (droplet) breakup depends on two dimensionless parameters: the Weber number, expressing the ratio of aerodynamic to surface tension forces, and the Ohnesorge number, expressing the ratio of viscous to surface tension forces. For example, depending on the combination of the above parameters, different types of secondary breakup occur, as illustrated in Figure 4, where a map of the secondary breakup regimes is presented. Figure 4 is based on the findings of [6], as compiled in [7]; here the areas of typical gasoline and diesel sprays for automotive applications are identified. Evidently, for an accurate modeling of spray breakup, a proper representation of the thermophysical properties affecting the values of Weber and Ohnesorge numbers (fuel viscosity, surface tension and density), is essential. Further, for accurately modeling fuel evaporation, accurate values of vapor pressure, latent heat of evaporation and fuel enthalpy are required. Hence, a detailed model of residual HFO thermophysical properties was developed in [8], and tested by CFD simulations of spray breakup in a constant volume combustion chamber. Figures 5, 6 and 7 present the dependence of dynamic viscosity, surface tension and vapor pressure on temperature, for HFO residual (and diesel) fuel, based on the development reported in [8]. As lighter components are also present in the HFO used in marine Diesel engines, multicomponent fuel models are needed for a more accurate description of the spray dynamics and fuel evaporation. To this end, following the development of [8], a two-component evaporation model was developed in [9]. The model assumes uniform droplet composition, with a residual and a lighter component, each of properly defined thermophysical properties. The new model was tested for spray breakup and combustion in constant volume combustion chambers (see Figure 8) [9]. Based on the computational tools developed, next steps in the research work of the NTUA team will include modeling and optimization of HFO combustion in large marine Diesel engines.

1 4 0

5 20 4

Table 2. Design variables and their prescribed ranges in the optimization study of [3].

Figure 2. Sketch of the generic injection profile with pilot injection utilized in the optimization studies of [3]. Three design variables (SOPI, SOMI and PMF) are indicated with arrows.

Figure 3. Final Pareto fronts (normalized values of the exhaust NOx concentration and SFOC) for the constrained and the unconstrained optimization studies reported in [3].

Breakup Regimes 100000

10000 Catastrophic 1000 Weber Shear 100 Multimode Bag 10 Osc. Deform. 1 0.001 0.01

Diesel

Gasoline

0.1 Ohnesorge

10

Figure 8. HFO injection in a constant volume chamber: color coded contours of temperature at different time instants. The droplets still in liquid phase are presented in white color [9].

Figure 4. Map of secondary atomization regimes as functions of Ohnesorge and Weber numbers, in which the areas representative of automotive gasoline and diesel sprays are identified [10].
100000

2.3 WATER ADDITION STUDIES A promising approach for substantial NOx reduction in marine Diesel engines is the introduction of water in the combustion chamber. Water can be introduced in the following ways: (a) air fumigation, in which water is added to the intake air, (b) Direct Water Injection (DWI) into the combustion chamber, and (c) fuel-water emulsions. The potential of water addition, in particular air fumigation and direct injection in a large two-stroke marine engine has been explored in the computational studies reported in [10], [11], in which the formation of NO has been modeled based on the extended Zeldovich mechanism [12]. Soot formation has been also taken into account by means of the Hiroyasu soot model [13]. The engine considered in [10], [11] is the large two-stroke marine Diesel engine also utilized in the optimization study of [3] (see Table 1). In the air fumigation simulations, a homogeneous mixture has been assumed at the beginning of compression. In the simulations corresponding to direct water injection, the position of injectors as well as the injection profile was initially assumed the same for both fuel and water. Figures 9 and 10 present the computed histories of NO concentration, normalized by the final concentration of the reference case (no water addition). Clearly, direct water injection has a more drastic effect on NO production. In both cases, the reduced levels of NO should be primarily attributed to the increased specific heat of the mixture, and to a smaller extent to the reduced amount of O2 available for combustion. Specific NO emissions can be defined in terms of the exhaust concentration divided by the work output of the closed part of the engine cycle. In Figure 11, the decrease in specific NO emissions is presented versus the increase in SFOC (in comparison to the reference design), for both water addition techniques. Evidently Direct Water Injection leads to significantly higher NO reduction, compared to the one achieved with air fumigation. In addition, reduced amounts of water are required with DWI.

Dynamic Viscosity [cP]

10000 1000 100 10 1 0.1 250

Diesel HFO

300

350

400

450

Temperature [K]

Figure 5. Dynamic viscosity for diesel fuel and HFO, as a function of temperature [8].
0.035 Diesel

Surface Tension [N/m]

0.03

HFO

0.025

0.02

0.015 250

300

350

400

450

Temperature [K]

Figure 6. Surface tension for diesel fuel and HFO, as a function of temperature [8].
1.00E+00 Vapor Pressure [bar] 1.00E-01 1.00E-02 1.00E-03 1.00E-04 1.00E-05 1.00E-06 1.00E-07 250 300 350 Temperature [K] 400 Diesel HFO

450

Figure 7: Vapor pressure for diesel fuel and HFO, as a function of temperature [8].
4


1.2 Reference 1 50% w ater 100% w ater 150% w ater 0.8 200% w ater

0.6

0.4

0.2

0 0 10 20 30 40 50 60

soot is lower with the DWI technique. The significant increase in soot emissions for large quantities of water directly injected into the combustion chamber is attributed to the shift of the overall mixture to conditions (temperature and equivalence ratio) that favor soot formation; this is further elaborated below. Moreover, the relatively lower temperatures hinder the soot oxidation, leading to larger soot quantities in the exhaust gases.
Crank Angle [deg]

Figure 9. Computed average NO concentration histories, normalized by the final value of the reference case, for different values of added water mass by means of air fumigation, and for the reference case [10].

1.2 Ref erence 1 20% w ater 40% w ater 50% w ater 0.8 60% w ater 80% w ater

[-]

O [-]

0.6

0.4

0.2

0 0 10 20 30 40 50 60

Crank Angle [deg]

Figure 12. Change in specific NO emissions, with respect to reference case, versus corresponding change in soot emissions, for different levels of water addition, for air fumigation and DWI techniques. The numbers adjacent to the curves indicate the percentage of water mass [11].

Figure 10. Computed average NO concentration histories, normalized by the final value of the reference case, for different values of directly injected water mass, and for the reference case [10].

Figure 11. Change in specific NO emissions, with respect to reference case, versus corresponding SFOC change, at different levels of water addition, for air fumigation and DWI techniques [10].

In CFD studies of Diesel engines, understanding of the emissions formation processes can be enhanced by utilizing temperature - equivalence ratio (T-) maps. T- maps present concentrations of produced NOx and soot, as functions of local stoichiometry and temperature. It is noted that for soot, the maps provide information on the soot formation process only. Here, the maps of Kitamura et al. [14] are used. Figure 13 presents T- maps including the local T, values of all computational cells, for a representative crank angle value (18o CA aTDC), for the reference case, and for two water addition cases, corresponding to 50% of air fumigation and DWI, respectively. Figure 13 demonstrates that water addition results in a shift from the NOx production to the soot production regime, which is very pronounced for the DWI case. In conjunction with the lower oxidation rates due to the lower temperature levels, this results in the computed higher exhaust soot levels in the case of DWI.

In Figure 12, the change in specific NO emissions is presented, versus the change in soot emissions, for both techniques. Clearly, soot can increase dramatically when DWI is used. However, for a given reduction of NO emissions, the increase in
5

3. SUMMARY
The paper has summarized recent research at NTUA, related to modeling and optimization of flow and combustion processes in marine Diesel engines. The models and software developed enable

the coupling of CFD codes with optimization tools, as well as an accurate representation of HFO thermophysical properties, which is essential for spray breakup and evaporation modeling. They also enable detailed computational water addition studies. Some of the results presented verify that a complex combination of in-cylinder measures, including water addition and possibly EGR, is necessary to meet the 2016 regulations (without exhaust gas aftertreatment). These optimized measures could be identified by CFD-based optimization studies. Based on the computational tools developed by the NTUA research team, such studies are feasible, and are planned for the near future.

4. ACKNOWLEDGMENTS
This work has been supported by the EU projects MARINECFD and MARINELIVE. The authors would like to thank Mr. P. Andreadis, Mr. A. Frangopoulos, Mr. N. Kyriakides, Mr. N. Stamoudis and Mr. A. Zompanakis, past and present NTUA students, for obtaining most of the results presented here.

REFERENCES
1. Service Contract on Ship Emissions: Assignment, Abatement and Market-based Instruments. Final Report for the European Commission, Directorate General Environment, Entec UK Ltd, Northwich, Cheshire, UK, August 2005. 2. MEPC.176(58) Amendments to the Annex of the Protocol of 1997 to amend the International Convention for the Prevention of Pollution from Ships, 1973, as modified by the Protocol of 1978 relating thereto (Revised MARPOL Annex VI), International Maritime Organization, London, UK, October 2008. 3. Andreadis, P., Zompanakis, A., Chryssakis, C., Kaiktsis, L. 2011, "Effects of Fuel Injection Parameters on the Performance and Emissions Formation in a Large-Bore Marine Diesel Engine", International Journal of Engine Research, Vol. 12, no 1, pp. 14-29. 4. Amsden, A.A. 1993, KIVA-3: a KIVA Program with Blockstructured Mesh for Complex Geometries, LA-12503-MS, Los Alamos National Laboratory, Los Alamos, New Mexico, USA. 5. Karakasis, M., Giannakoglou, K.C. 2006, On the Use of Metamodel-assisted Multi-Objective Evolutionary Algorithms, Engineering Optimization, Vol. 38, no 8, pp. 941957. 6. Faeth, G.M., Hsiang, L.P., Wu, P.K. 1995, Structure and Breakup Properties of Sprays, International Journal of Multiphase Flow, Vol. 21, Suppl., pp. 99-127. 7. Chryssakis, C., Assanis, D.N. 2008, A Unified Fuel Spray Breakup Model for Internal Combustion Engine Applications, Atomization and Sprays, Vol. 18, no 5, pp. 375-426. 8. Kyriakides, N., Chryssakis, C., Kaiktsis, L. 2009, Influence of Heavy Fuel Properties on Spray Atomization for Marine Diesel Engine Applications, SAE Technical Paper Series 2009-01-1858, SAE International Powertrains, Fuels & Lubricants Meeting, Florence, Italy. 9. Stamoudis, N., Chryssakis, C. Kaiktsis, L. 2011, A Two-Component Heavy Fuel Oil Evaporation Model for CFD Studies in Marine Diesel Engines, Fuel, under review.

Figure 13. T- maps, including local T, values of all computational cells, at 18o CA aTDC, for the reference case, and air fumigation and DWI with 50% water.

10. Frangopoulos, A., Chryssakis, C., Kaiktsis, L. 2010, Computational Investigation of InCylinder NOx Emissions Reduction in a Large Marine Diesel Engine Using Water Addition Strategies, Proceedings, Paper 2010-01-1257, SAE World Congress 2010, Detroit, USA. 11. Frangopoulos, A., Chryssakis, C., Kaiktsis, L. 2010, Computational Study of In-Cylinder NOx Reduction in a Large Marine Diesel Engine Using Water Injection Strategies, Proceedings, Paper No. 229, CIMAC Congress 2010, Bergen, Norway. 12. Heywood, J.B. 1988, Internal Combustion Engine Fundamentals, McGraw-Hill, 1988. 13. Hiroyasu, H., Kadota, T., Arai, M. 1983, Development and Use of a Spray Combustion Modeling to Predict Diesel Engine Efficiency and Pollutant Emissions (Part I: Combustion Modeling), Bulleting of the JSME, Vol. 26, pp. 569-575. 14. Kitamura, T., Ito, T., Send, J., Fujomoto, H. 2002, Mechanism of Smokeless Diesel Combustion with Oxygenated Fuels Based on the Dependence of the Equivalence Ratio and Temperature on Soot Particle Formation", International Journal of Engine Research, Vol. 3, no 4, pp. 223-247.

Vous aimerez peut-être aussi