Next Article in Journal
Live Music in the Time of Corona: On the Resilience and Impact of a Philharmonic Orchestra on the Urban Economy
Next Article in Special Issue
A Migration Learning Method Based on Adaptive Batch Normalization Improved Rotating Machinery Fault Diagnosis
Previous Article in Journal
Grit and Career Construction among Chinese High School Students: The Serial Mediating Effect of Hope and Career Adaptability
Previous Article in Special Issue
Assessment of an Optimal Design Method for a High-Energy Ultrasonic Igniter Based on Multi-Objective Robustness Optimization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Computational Investigation of Combustion, Performance, and Emissions of a Diesel-Hydrogen Dual-Fuel Engine

1
CRRC Academy Corporation Limited, Beijing 100160, China
2
School of Mechanical Engineering, Beijing Institute of Technology, Beijing 100081, China
3
Beijing Lab of New Energy Vehicles and Key Lab of Regional Air Pollution Control, College of Energy and Power Engineering, Beijing University of Technology, Beijing 100124, China
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(4), 3610; https://doi.org/10.3390/su15043610
Submission received: 23 January 2023 / Revised: 9 February 2023 / Accepted: 13 February 2023 / Published: 15 February 2023

Abstract

:
This paper aims to expose the effect of hydrogen on the combustion, performance, and emissions of a high-speed diesel engine. For this purpose, a three-dimensional dynamic simulation model was developed using a reasonable turbulence model, and a simplified reaction kinetic mechanism was chosen based on experimental data. The results show that in the hydrogen enrichment conditions, hydrogen causes complete combustion of diesel fuel and results in a 17.7% increase in work capacity. However, the increase in combustion temperature resulted in higher NOx emissions. In the hydrogen substitution condition, the combustion phases are significantly earlier with the increased hydrogen substitution ratio (HSR), which is not conducive to power output. However, when the HSR is 30%, the CO, soot, and THC reach near-zero emissions. The effect of the injection timing is also studied at an HSR of 90%. When delayed by 10°, IMEP improves by 3.4% compared with diesel mode and 2.4% compared with dual-fuel mode. The NOx is reduced by 53% compared with the original dual-fuel mode. This study provides theoretical guidance for the application of hydrogen in rail transportation.

1. Introduction

Adjusting the industrial and energy structures is inevitable to realize an emission peak and carbon neutrality [1,2]. The gaseous and particle emissions from internal combustion engines (ICEs) contribute an important proportion of total atmospheric pollutants [3,4]. Therefore, searching for low-carbon or zero-carbon fuels has become an important research direction for developing ICEs [5,6,7]. Hydrogen is regarded as the clean energy with the most development potential in the 21st century because of many advantages, such as diverse sources, being clean and low-carbon, being flexibile and efficient, and having various application scenarios [8,9,10]. Hydrogen energy has become the preferred direction for the new round of carbon emission reduction and carbon neutrality worldwide [11,12]. It has been incorporated into the energy strategy deployment by many countries [13,14]. From the strategic point of view of energy security and sustainable development, China has considered hydrogen energy a new strategic industry for development [15,16].
There are two main ways to utilize hydrogen energy: fuel cells [17] and ICEs [18]. Fuel cells have the advantages of high efficiency and zero-emissions, but they are technically complex, costly, and dependent on the construction of supporting systems [19]. The hydrogen-fueled ICEs can use industrial by-product hydrogen to convert energy by the combustion method to achieve similar thermal efficiency as fuel cells, which has the significant advantage of low cost [20]. Hydrogen-fueled ICEs retain the main structure and system of traditional ICEs [21,22,23]. Based on traditional ICEs, hydrogen-fueled ICEs can be realized by simply replacing the hydrogen supply and injection system, hydrogen-specific cold spark plugs, matching a new turbocharger, and adapting the lubrication and crankcase ventilation accordingly [24,25,26]. Therefore, hydrogen-fueled ICEs are an essential technology direction to promote the upgrading and transformation of various application fields of traditional ICEs, which help achieve peak emissions and carbon neutrality [27,28,29,30].
For the spark-ignited hydrogen-fueled ICEs, the higher laminar flame speed and larger dilute combustion limits of hydrogen allow the engine to operate over a wide range of equivalence ratios, in which the thermal efficiency typically equals or exceeds that of gasoline-fueled engines [31]. It has been shown that spark-ignited hydrogen-fueled ICEs can achieve near-zero NOx emissions under lean-burn operating conditions [32]. However, hydrogen-fueled ICEs face problems such as backfire, premature ignition, and detonation at high loads, which severely limit their rapid development [33]. The heat transfer loss through the wall increases rapidly as the equivalent ratio increases [34]. The problems of detonation and premature ignition are challenging for heavy engines, which has prompted research on hydrogen compression ignition (CI) engines [35]. Therefore, dual-fuel technology was developed for engines [36]. The diesel fuel is designed to assist the ignition of the hydrogen fuel, known as the diesel pilot ignition mode [37,38]. In fact, this concept has long been widely used in natural gas-fueled ICEs [39] and extensively used in the marine field [40,41]. Tripathi et al. [42] investigated the performance and emissions of a diesel-hydrogen dual-fuel engine using numerical simulations. The results showed that the combination of two injection strategies can simultaneously reduce NOx emissions and improve IMEP. Sharma et al. [43] investigated the performance and emissions of dual-fuel engines with different compression ratios and hydrogen fractions through a similar methodology. The results show that compression-ignition mode is not suitable for compression ratios less than 14.5. Köse et al. [44] experimentally investigated that hydrogen enrichment reduced pollutant emissions except for NOx and increased brake thermal efficiency (BTE) and exhaust temperature. When operated at 1750 rpm, 40.4% BTE was achieved at 2.5% hydrogen enrichment, while in diesel mode, the BTE was 33%. Ramsay et al. [45] studied the effect of the constant volume combustion phase on the performance and emissions of a dual-fuel engine under various load and hydrogen energy share conditions. The results demonstrated that this method could improve thermal efficiency with far lower carbon-based emissions under all conditions. Taghavifar et al. [46], through a 1-D model, parametrically investigated the effects of levels of diesel and hydrogen, compressor pressure ratio, and combustion duration on energy, exergy, and performance in a diesel-hydrogen dual-fuel engine. The results indicated that supercharging can significantly improve thermal efficiency and reduce fuel consumption. Wu et al. [47] optimized the operation parameters of a dual-fuel engine based on the Taguchi method. The results revealed that for NOx, using EGR technology reduces more than 60.5% at various loads, and BSFC can reduce it by 14.52%.
From the above literature, it is clear that hydrogen can significantly improve thermal efficiency and performance. It is essential to improve thermal efficiency and reduce carbon emissions for rail transportation. Therefore, in this paper, hydrogen enrichment and hydrogen substitution are investigated separately by using numerical simulation. In addition, injection timing studies are carried out to optimize performance and emissions. This paper explored the potential of hydrogen in ICEs for rail transportation and provided a theoretical basis for practical applications.

2. Materials and Methods

2.1. Numerical Methodology

In this work, the prototype engine is a high-speed diesel engine with a total displacement of 87.54 L. The cylinder bore, stroke, and compression ratio were 180 mm, 215 mm, and 17, respectively. The engine is designed to meet future emissions regulations by taking into account compactness, power-to-weight ratio, economy, and reliability. The CONVERGE code was applied to calculate the flow motion and combustion phenomena in the combustion chamber [48,49,50]. The SolidWorks software was used to establish the 3-D geometric model. The model (*.stl) was then imported into CONVERGE to calculate the combustion. To simulate the turbulence, spray, and combustion, the mathematical models adopted in CFD calculations are summarized in Table 1 [51]. The dual-fuel reaction mechanism with 76 species and 464 reactions was chosen to simulate the combustion of the hydrogen and diesel mixture [52]. This dual-fuel reaction mechanism was coupled with the GRI3.0 mechanism, which can accurately simulate hydrogen combustion. This mechanism was used in the simulation of pure diesel and dual-fuel, which avoids the influence of mechanism differences on the results [53]. In addition, at 1800 rpm, the temperatures of the piston, cylinder wall, and cylinder head were set to 553, 433, and 523 K, respectively.

2.2. Model Validation

To reduce calculation time, only 1/8 of the domain was recorded and analyzed since the injection is located in the center of the chamber with eight nozzles. As illustrated in Figure 1a, the cylindrical domain was classified into angular sectors such that one injector falls at the center of each sector. Figure 1b shows the computational mesh at TDC, in which the adaptive mesh refinement and fixed embedding were activated to guarantee the calculation accuracy. Figure 2 shows the predicted pressure profile under different meshes. It can be seen from Figure 2 that the 4 mm basic grid can meet the calculation accuracy. As the dual-fuel mode is still in the development stage, only the diesel mode was tested. To validate the combustion and turbulence model, the model was verified at speeds of 600, 1400, and 1800 rpm. Figure 3 shows the comparison between experimental and simulated cylinder pressure. The calculation accuracy was high and met the engineering requirements for the next step of research. Since the reaction kinetics mechanism used includes hydrogen and diesel, the verified model was used for the next study. In addition, Tripathi et al. [42] was validated using the same turbulence model and mechanism in diesel-hydrogen conditions, which is another indication that the accuracy of the model in this paper can be studied in the next part [54].

2.3. Research Schemes

The present work aims to evaluate the potential of hydrogen in diesel engines and to explore the maximum hydrogen substitution ratio for the purpose of reducing carbon emissions. In all the simulations, the speed was kept at 1800 rpms, and the total ejected energy was same as with the 18-bar BMEP. In the hydrogen enrichment test, the source diesel was kept the same, and the hydrogen was the added energy. The hydrogen enrichment ratio ( H E R ) was between 10 and 20%. In the hydrogen substitution test, the total energy was kept the same as with the origin diesel, and the substitution ratios ( H S R ) were 30, 60, and 90%, respectively. In the relative injection timing test, the high H S R was tested with varing injections. The summary of the test conditions is listed in Table 2.
The definition of H E R and H S R is as follows:
H E R = m H 2 LHV H 2 m Diesel LHV Diesel
H S R = m H 2 LHV H 2 m Diesel LHV Diesel + m H 2 LHV H 2

3. Results and Discussion

3.1. Effect of Hydrogen Enrichment

The in-cylinder pressure is an important parameter to deeply understand the combustion process and directly affects performance and emissions. The comparison of in-cylinder pressures is shown in Figure 4. In this section, D and H represent diesel and hydrogen, respectively. D100 means the diesel energy fraction is 100%, and H10 represents the hydrogen energy fraction at 10%. As shown in Figure 4, a larger H E R leads to a higher in-cylinder pressure [42]. The corresponding CA position (PFP_CA) is delayed with the increasing H E R , as shown in Table 3. The IMEP of D100 + H10 and D100 + H20 increased by 8.7 and 17.7%, respectively. The temperature contour is shown in Table 4. It is obvious that the temperature increases with the H E R . This is because diesel has a dominant effect on combustion, while hydrogen only promotes combustion. The comparison of HRR and combustion phases is shown in Figure 5. The profiles of HRR and combustion phases are similar to each other. The phases are delayed with increasing H E R , but the level is minor. This is because the overall energy of the additional hydrogen enrichment condition is higher than the original engine. Although the hydrogen enrichment could enhance the flame speed, the ratio is relatively small.
The comparison of emissions is shown (Figure 6) at hydrogen additional enrichment, where the value is recorded at the exhaust opening timing. The data are scaled so that it can be displayed on a single figure. In general, all pollutant emissions except NOx and CO2 decrease with the increase of H E R . This is because soot, THC, and CO can all be further oxidized and burn more completely as the H E R increases. The higher H A R results in a higher in-cylinder temperature, which is mainly responsible for the complete oxidation of C O + O H C O 2 + H and C O + O + M C O 2 + M , and improving the degree of complete combustion consequently [55]. The NOx formation is chiefly determined by mean temperature. For the case D100 + H2, the NOx emissions increased by 45%. It is recommended to combine after-treatment equipment for further optimization.

3.2. Effect of Hydrogen Substitution

In order to investigate the effect of hydrogen substitution enrichment on combustion, performance, and emissions, the conditions with H S R of 0, 30, 60, and 90% are selected for further simulation. In all the simulations, the total energy is kept the same, where D and H represent diesel and hydrogen, respectively. The D10-H90 means the diesel energy fraction is 10% and the hydrogen energy fraction is 90%. The in-cylinder pressure is shown in Figure 7. The comparison of PFP, corresponding CA, and IMEP is listed in Table 5. The PFP increases with the increase of H S R . As the H S R increases, the corresponding CA advances. This is because the higher burning velocity of hydrogen accelerates the flame speed and results in a higher PFP and an advanced CA. The IMEP of D10-H90 is lower than D40-H60, although the PFP is higher. However, too high a PFP is also not conducive to the modification of the original engine and even requires a redesign of the overall strength.
As shown in Table 6, the mean temperature is raised with the increase of H S R . The high-temperature zone appears earlier as the H S R increases. When the mixing ratio is 90%, the high-temperature zone appears inside the combustion chamber, while the high-temperature zone at other ratios is beyond the combustion chamber near the cylinder head. Therefore, increasing H S R is beneficial to reduce the heat load on the cylinder head. The high temperature zone inside the cylinder is more uniform with the increase of H S R . This is due to the fact that as the H S R increases, the combustion mode gradually shifts from diffusion combustion to premixed combustion mode. This transition can be visualized from the HRR, as shown in Figure 8a. As the H S R increases, the HRR gradually shows a dual trend. The first peak is caused by diesel igniting hydrogen, and the second peak is caused by the simultaneous ignition of hydrogen at multiple points of the combustion flame. When the H S R is 60%, the two values are similar, while the peak of hydrogen combustion exotherm is much higher than that of diesel when the H S R is 90%. It can also be seen from Table 6 that diffusion combustion generally starts from the tail, and the high temperature area is small. Because the laminar flame of hydrogen is faster, when premixed, the flame quickly spreads throughout the combustion chamber. Therefore, when the H S R is higher, the second peak of heat release is also higher. The comparison of combustion phases is also shown in Figure 8b. Due to the faster burning rate of hydrogen and the shift in the combustion mode, the combustion phases are advanced. However, when the H S R is 90%, the CA50 is already before the TDC, which is not conducive to power output.
For the emissions, as shown in Figure 9, all pollutant emissions except NOx are reduced, especially THC and CO, which have become small in order of magnitude at H S R >30%. It can be analyzed in two ways: first, the increased H S R reduces the diesel fuel, resulting in lower carbon for THC and CO, i.e., the total fuel carbon content. Secondly, due to the higher combustion temperature of hydrogen enrichment, hydrogen increases the free radical content of the reaction, which promotes the combustion process and makes the combustion more adequate. For soot, when the H S R is higher, it leads to more homogeneous fuel mixing due to the lower mixture density. In addition, the hydrogen raises the temperature of the compression process, resulting in a lower in-cylinder temperature gradient, which deteriorates the soot generation environment. The increase in combustion temperature provides a good environment for soot oxidation. The synergistic effect of the two leads to extremely low soot. For CO2, because hydrogen is a carbon-free fuel, CO2 emissions are dramatically reduced at higher H S R [56].

3.3. Effect of Pilot Injection Timing

From the above study, it can be concluded that hydrogen enrichment improves the combustion process and reduces pollutant emissions except NOx. For performance, the optimization of the injection strategy is necessary because the combustion phase at high H S R conditions is too advanced [57]. In addition, at high H S R , the PFP is too high, resulting in excessive mechanical load. In addition, considering the goal of carbon reduction and the common operating conditions of rail internal combustion engines, this section investigates the effect of the pilot fuel injection timing on the performance, combustion, and emissions of dual-fuel engines.
The “I” represents the relative injection timing, i.e., H0-I0 is the injection timing of the original engine in diesel mode (the relative injection timing is 0 °CA). The H90-I5 represents 90% hydrogen energy and 10% diesel energy, and the diesel injection timing is delayed by 5 °CA. Figure 10 shows the comparison of in-cylinder pressure at different injection timings. In all cases, the PFP is higher than the diesel mode. The pressure rise rate is also increased in dual-fuel mode. As listed in Table 7, the CA position of PFP decreases with the delayed injection timing. According to the above study, in dual-fuel mode, when at the maximum H S R , the delayed injection timing facilitates an increase in engine power output. The IMEP exceeds the original diesel condition. When delayed by 10 °CA, IMEP improves by 3.4% compared with diesel mode and 2.4% compared with dual-fuel mode.
The contour of temperature distribution is illustrated in Table 8. The appearance of the high temperature zone is delayed with the delay in injection timing. The mean temperature in the dual-fuel mode is higher than the diesel model, which undoubtedly causes higher NOx emissions. The trend of peak temperature is the same as the injection timing. When the injection timing is early, two separate high-temperature zones appear at the beginning of combustion. This is due to lower temperatures and incomplete fuel atomization. Therefore, the fuel in the combustion chamber and near the injector ignited hydrogen gas separately. When in the dual-fuel mode, the diesel is already atomized and broken at the end of the injection compared with the pure-diesel mode. This is due to the reduced in-cylinder ambient density, which facilitates fuel atomization. The HRR and combustion phases are shown in Figure 11. The combustion center is advanced with the early injection timing. This is because as the injection timing is delayed, the in-cylinder pressure and temperature decrease, which delay the combustion phases [58]. As the injection timing is delayed, the second peak of heat release rises initially before decreasing. When the injection timing is delayed by 15 °CA, the second peak of heat release dissipates, resulting in a shift to premixed combustion.
The effect of the injection timing on the emissions is shown in Figure 12. The soot, THC, CO, and CO2 are much smaller than the original engine after hydrogen enrichment conditions. Alternatively, the carbon content of the fuel is reduced to 10% of the original engine because the pilot diesel is 1/10 of the original engine. For CO2, it is reduced by nearly 90%. Conversely, due to the higher combustion temperature in dual-fuel mode, it leads to a dramatic reduction in the production of unburned emissions. At high H S R , the lower in-cylinder temperature gradient due to less diesel injection mass and more homogeneous mixing is not conducive to soot production. In addition, due to the higher temperature, a good environment for soot oxidation is provided. When the injection timing was delayed by 15 °CA, a small amount of THC appeared. For NOx generation, the NOx generation decreases with the delayed injection timing. This is because the delayed injection timing reduces the in-cylinder combustion temperature, which is not conducive to NOx generation [42]. Compared with the original injection timing, NOx is reduced by 53% when delaying 10 °CA.

4. Conclusions

In this paper, the effect of hydrogen on diesel combustion is investigated using a numerical method. The effects of additional hydrogen enrichment, hydrogen substitution, and pilot injection timing on combustion, performance, and emissions are investigated. The main findings are as follows:
(1)
The hydrogen enrichment was used for the study at very high loads. With the increase of H E R , the diesel fuel atomized better and burned more fully. When in a small H E R , the combustion phase of the engine had a small range of variation and a consistent shape of the HRR. Due to the increase in temperature after hydrogen enrichment, it leads to higher NOx. However, CO, soot, and THC emissions were reduced due to more complete combustion.
(2)
The hydrogen substitution was also studied in order to reduce carbon emissions and increase the H S R . As H S R increases, the peak cylinder pressure increases, and the combustion phase advances. The higher combustion temperature of hydrogen leads to more NOx. When the H S R was 90%, the center of combustion was located unfavorably in front of the upper stop, resulting in a lower IMEP.
(3)
Since the center of gravity of combustion is too advanced at a higher H S R , the study of injection timing was carried out. With the delay in injection timing, the in-cylinder pressure decreases and the combustion temperature decreases. In addition, the work capacity increases and the NOx decreases.
(4)
The implementation of hydrogen injectors necessitates increased control standards, as the low energy density of hydrogen may cause limitations in larger-scale applications. In addition, hydrogen safety must be taken into account.
In conclusion, this paper investigates the effect of hydrogen on diesel engine combustion and emissions and provides theoretical guidance for practical design optimization. Subsequent work will couple EGR with an injection strategy to further increase power and reduce NOx emissions.

Author Contributions

Formal analysis, B.Z. and H.W.; Funding acquisition, S.W.; Investigation, H.W.; Methodology, B.Z.; Project administration, S.W.; Resources, S.W.; Software, B.Z. and H.W.; Supervision, S.W.; Writing—original draft, B.Z.; Writing—review & editing, H.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by CRRC Academy Corporation Limited and the BIT Research and Innovation Promoting Project (Grant No.2022YCXZ004). The authors also acknowledge all students involved in tests and calculations for this work.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy restrictions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gong, C.; Li, Z.; Sun, J.; Liu, F. Evaluation on combustion and lean-burn limit of a medium compression ratio hydrogen/methanol dual-injection spark-ignition engine under methanol late-injection. Appl. Energy 2020, 277, 115622. [Google Scholar] [CrossRef]
  2. Bakar, R.A.; Widudo; Kadirgama, K.; Ramasamy, D.; Yusaf, T.; Kamarulzaman, M.K.; Sivaraos; Aslfattahi, N.; Samylingam, L.; Alwayzy, S.H. Experimental analysis on the performance, combustion/emission characteristics of a DI diesel engine using hydrogen in dual fuel mode. Int. J. Hydrog. Energy 2022. [Google Scholar] [CrossRef]
  3. Gong, C.; Si, X.; Liu, F. Effects of injection timing and CO2 dilution on combustion and emissions behaviors of a stoichiometric GDI engine under medium load conditions. Fuel 2021, 303, 121262. [Google Scholar] [CrossRef]
  4. Wang, H.; Ji, C.; Shi, C.; Ge, Y.; Meng, H.; Yang, J.; Chang, K.; Wang, S. Comparison and evaluation of advanced machine learning methods for performance and emissions prediction of a gasoline Wankel rotary engine. Energy 2022, 248, 123611. [Google Scholar] [CrossRef]
  5. Gong, C.; Yi, L.; Zhang, Z.; Sun, J.; Liu, F. Assessment of ultra-lean burn characteristics for a stratified-charge direct-injection spark-ignition methanol engine under different high compression ratios. Appl. Energy 2020, 261, 114478. [Google Scholar] [CrossRef]
  6. Duan, X.; Xu, L.; Jiang, P.; Lai, M.-C.; Sun, Z. Parallel perturbation analysis of combustion cycle-to-cycle variations and emissions characteristics in a natural gas spark ignition engine with comparison to consecutive cycle method. Chemosphere 2022, 308, 136334. [Google Scholar] [CrossRef]
  7. Li, Y.; Gao, W.; Zhang, P.; Ye, Y.; Wei, Z. Effects study of injection strategies on hydrogen-air formation and performance of hydrogen direct injection internal combustion engine. Int. J. Hydrog. Energy 2019, 44, 26000–26011. [Google Scholar] [CrossRef]
  8. Ma, Y.; Wang, X.; Li, T.; Zhang, J.; Gao, J.; Sun, Z. Hydrogen and ethanol: Production, storage, and transportation. Int. J. Hydrog. Energy 2021, 46, 27330–27348. [Google Scholar] [CrossRef]
  9. Sun, X.; Liu, H.; Duan, X.; Guo, H.; Li, Y.; Qiao, J.; Liu, Q.; Liu, J. Effect of hydrogen enrichment on the flame propagation, emissions formation and energy balance of the natural gas spark ignition engine. Fuel 2022, 307, 121843. [Google Scholar] [CrossRef]
  10. Seelam, N.; Gugulothu, S.K.; Reddy, R.V.; Bhasker, B.; Kumar Panda, J. Exploration of engine characteristics in a CRDI diesel engine enriched with hydrogen in dual fuel mode using toroidal combustion chamber. Int. J. Hydrog. Energy 2022, 47, 13157–13167. [Google Scholar] [CrossRef]
  11. Liu, W.; Wan, Y.; Xiong, Y.; Gao, P. Green hydrogen standard in China: Standard and evaluation of low-carbon hydrogen, clean hydrogen, and renewable hydrogen. Int. J. Hydrog. Energy 2022, 47, 24584–24591. [Google Scholar] [CrossRef]
  12. Yu, X.; Li, G.; Du, Y.; Guo, Z.; Shang, Z.; He, F.; Shen, Q.; Li, D.; Li, Y. A comparative study on effects of homogeneous or stratified hydrogen on combustion and emissions of a gasoline/hydrogen SI engine. Int. J. Hydrog. Energy 2019, 44, 25974–25984. [Google Scholar] [CrossRef]
  13. Duan, X.; Liu, Y.; Liu, J.; Lai, M.-C.; Jansons, M.; Guo, G.; Zhang, S.; Tang, Q. Experimental and numerical investigation of the effects of low-pressure, high-pressure and internal EGR configurations on the performance, combustion and emission characteristics in a hydrogen-enriched heavy-duty lean-burn natural gas SI engine. Energy Convers. Manag. 2019, 195, 1319–1333. [Google Scholar] [CrossRef]
  14. Arat, H.T. Alternative fuelled hybrid electric vehicle (AF-HEV) with hydrogen enriched internal combustion engine. Int. J. Hydrog. Energy 2019, 44, 19005–19016. [Google Scholar] [CrossRef]
  15. Shi, X.; Zheng, Y.; Lei, Y.; Xue, W.; Yan, G.; Liu, X.; Cai, B.; Tong, D.; Wang, J. Air quality benefits of achieving carbon neutrality in China. Sci. Total Environ. 2021, 795, 148784. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, R.; Hanaoka, T. Deployment of electric vehicles in China to meet the carbon neutral target by 2060: Provincial disparities in energy systems, CO2 emissions, and cost effectiveness. Resour. Conserv. Recycl. 2021, 170, 105622. [Google Scholar] [CrossRef]
  17. Li, J.; Liang, M.; Cheng, W.; Wang, S. Life cycle cost of conventional, battery electric, and fuel cell electric vehicles considering traffic and environmental policies in China. Int. J. Hydrog. Energy 2021, 46, 9553–9566. [Google Scholar] [CrossRef]
  18. Gao, J.; Tian, G.; Ma, C.; Balasubramanian, D.; Xing, S.; Jenner, P. Numerical investigations of combustion and emissions characteristics of a novel small scale opposed rotary piston engine fuelled with hydrogen at wide open throttle and stoichiometric conditions. Energy Convers. Manag. 2020, 221, 113178. [Google Scholar] [CrossRef]
  19. Hames, Y.; Kaya, K.; Baltacioglu, E.; Turksoy, A. Analysis of the control strategies for fuel saving in the hydrogen fuel cell vehicles. Int. J. Hydrog. Energy 2018, 43, 10810–10821. [Google Scholar] [CrossRef]
  20. Duan, X.; Lai, M.; Jansons, M.; Guo, G.; Liu, J. A review of controlling strategies of the ignition timing and combustion phase in homogeneous charge compression ignition (HCCI) engine. Fuel 2021, 285, 119142. [Google Scholar] [CrossRef]
  21. Luo, Q.; Sun, B. Experiments on the effect of engine speed, load, equivalence ratio, spark timing and coolant temperature on the energy balance of a turbocharged hydrogen engine. Energy Convers. Manag. 2018, 162, 1–12. [Google Scholar] [CrossRef]
  22. Shi, C.; Chai, S.; Wang, H.; Ji, C.; Ge, Y.; Di, L. An insight into direct water injection applied on the hydrogen-enriched rotary engine. Fuel 2023, 339, 127352. [Google Scholar] [CrossRef]
  23. Sun, B.; Luo, Q.; Bao, L. Development and trends of direct injection hydrogen internal combustion engine technology. J. Automot. Saf. Energy 2021, 12, 265–278. [Google Scholar] [CrossRef]
  24. Wang, H.; Ji, C.; Shi, C.; Yang, J.; Ge, Y.; Wang, S.; Meng, H.; Chang, K.; Wang, X. Parametric modeling and optimization of the intake and exhaust phases of a hydrogen Wankel rotary engine using parallel computing optimization platform. Fuel 2022, 324 Pt A, 124381. [Google Scholar] [CrossRef]
  25. Wang, H.; Ji, C.; Shi, C.; Ge, Y.; Meng, H.; Yang, J.; Chang, K.; Yang, Z.; Wang, S.; Wang, X. Modeling and parametric study of the performance-emissions trade-off of a hydrogen Wankel rotary engine. Fuel 2022, 318, 123662. [Google Scholar] [CrossRef]
  26. Wang, H.; Ji, C.; Su, T.; Shi, C.; Ge, Y.; Yang, J.; Wang, S. Comparison and implementation of machine learning models for predicting the combustion phases of hydrogen-enriched Wankel rotary engines. Fuel 2022, 310 Pt B, 122371. [Google Scholar] [CrossRef]
  27. Luo, Q.; Hu, J.; Sun, B.; Liu, F.; Wang, X.; Li, C.; Bao, L. Experimental investigation of combustion characteristics and NOx emission of a turbocharged hydrogen internal combustion engine. Int. J. Hydrog. Energy 2019, 44, 5573–5584. [Google Scholar] [CrossRef]
  28. Takagi, Y.; Mori, H.; Mihara, Y.; Kawahara, N.; Tomita, E. Improvement of thermal efficiency and reduction of NOx emissions by burning a controlled jet plume in high-pressure direct-injection hydrogen engines. Int. J. Hydrog. Energy 2017, 42, 26114–26122. [Google Scholar] [CrossRef]
  29. Takagi, Y.; Oikawa, M.; Sato, R.; Kojiya, Y.; Mihara, Y. Near-zero emissions with high thermal efficiency realized by optimizing jet plume location relative to combustion chamber wall, jet geometry and injection timing in a direct-injection hydrogen engine. Int. J. Hydrog. Energy 2019, 44, 9456–9465. [Google Scholar] [CrossRef]
  30. Wang, H.; Ji, C.; Shi, C.; Yang, J.; Wang, S.; Ge, Y.; Chang, K.; Meng, H.; Wang, X. Multi-objective optimization of a hydrogen-fueled Wankel rotary engine based on machine learning and genetic algorithm. Energy 2023, 263, 125961. [Google Scholar] [CrossRef]
  31. Ma, D.; Sun, Z. Progress on the studies about NOx emission in PFI-H2ICE. Int. J. Hydrog. Energy 2020, 45, 10580–10591. [Google Scholar] [CrossRef]
  32. Xu, P.; Ji, C.; Wang, S.; Cong, X.; Ma, Z.; Tang, C.; Meng, H.; Shi, C. Effects of direct water injection on engine performance in a hydrogen (H2)-fueled engine at varied amounts of injected water and water injection timing. Int. J. Hydrog. Energy 2020, 45, 13523–13534. [Google Scholar] [CrossRef]
  33. Szwaja, S. Dilution of fresh charge for reducing combustion knock in the internal combustion engine fueled with hydrogen rich gases. Int. J. Hydrog. Energy 2019, 44, 19017–19025. [Google Scholar] [CrossRef]
  34. Szwaja, S.; Naber, J.D. Dual nature of hydrogen combustion knock. Int. J. Hydrog. Energy 2013, 38, 12489–12496. [Google Scholar] [CrossRef]
  35. Evans, A.; Wang, Y.; Wehrfritz, A.; Srna, A.; Hawkes, E.; Liu, X.; Kook, S.; Chan, Q.N. Mechanisms of NOx production and heat loss in a dual-fuel hydrogen compression ignition engine. In Proceedings of the SAE Technical Paper; No. 2021-01-0527; SAE International: Warrendale, PA, USA, 2021. [Google Scholar]
  36. Babayev, R.; Andersson, A.; Dalmau, A.S.; Im, H.G.; Johansson, B. Computational characterization of hydrogen direct injection and nonpremixed combustion in a compression-ignition engine. Int. J. Hydrog. Energy 2021, 46, 18678–18696. [Google Scholar] [CrossRef]
  37. Wang, Y.; Evans, A.; Srna, A.; Wehrfritz, A.; Hawkes, E.; Liu, X.; Kook, S.; Chan, Q.N. A numerical investigation of mixture formation and combustion characteristics of a hydrogen-diesel dual direct injection engine. In Proceedings of the SAE Technical Paper; No. 2021-01-0526; SAE International: Warrendale, PA, USA, 2021. [Google Scholar]
  38. Liu, X.; Srna, A.; Yip, H.L.; Kook, S.; Chan, Q.N.; Hawkes, E.R. Performance and emissions of hydrogen-diesel dual direct injection (H2DDI) in a single-cylinder compression-ignition engine. Int. J. Hydrog. Energy 2021, 46, 1302–1314. [Google Scholar] [CrossRef]
  39. Chen, Z.; He, J.; Chen, H.; Geng, L.; Zhang, P. Comparative study on the combustion and emissions of dual-fuel common rail engines fueled with diesel/methanol, diesel/ethanol, and diesel/n-butanol. Fuel 2021, 304, 121360. [Google Scholar] [CrossRef]
  40. Mavrelos, C.; Theotokatos, G. Numerical investigation of a premixed combustion large marine two-stroke dual fuel engine for optimising engine settings via parametric runs. Energy Convers. Manag. 2018, 160, 48–59. [Google Scholar] [CrossRef] [Green Version]
  41. Stoumpos, S.; Theotokatos, G.; Boulougouris, E.; Vassalos, D.; Lazakis, I.; Livanos, G. Marine dual fuel engine modelling and parametric investigation of engine settings effect on performance-emissions trade-offs. Ocean. Eng. 2018, 157, 376–386. [Google Scholar] [CrossRef] [Green Version]
  42. Tripathi, G.; Sharma, P.; Dhar, A.; Sadiki, A. Computational investigation of diesel injection strategies in hydrogen-diesel dual fuel engine. Sustain. Energy Technol. Assess. 2019, 36, 100543. [Google Scholar] [CrossRef]
  43. Sharma, P.; Dhar, A. Compression ratio influence on combustion and emissions characteristic of hydrogen diesel dual fuel CI engine: Numerical Study. Fuel 2018, 222, 852–858. [Google Scholar] [CrossRef]
  44. Köse, H.; Ciniviz, M. An experimental investigation of effect on diesel engine performance and exhaust emissions of addition at dual fuel mode of hydrogen. Fuel Process. Technol. 2013, 114, 26–34. [Google Scholar] [CrossRef]
  45. Ramsay, C.J.; Dinesh, K.K.J.R.; Fairney, W.; Vaughan, N. A numerical study on the effects of constant volume combustion phase on performance and emissions characteristics of a diesel-hydrogen dual-fuel engine. Int. J. Hydrog. Energy 2020, 45, 32598–32618. [Google Scholar] [CrossRef]
  46. Taghavifar, H.; Nemati, A.; Salvador, F.J.; De la Morena, J. 1D energy, exergy, and performance assessment of turbocharged diesel/hydrogen RCCI engine at different levels of diesel, hydrogen, compressor pressure ratio, and combustion duration. Int. J. Hydrog. Energy 2021, 46, 22180–22194. [Google Scholar] [CrossRef]
  47. Wu, H.-W.; Wu, Z.-Y. Combustion characteristics and optimal factors determination with Taguchi method for diesel engines port-injecting hydrogen. Energy 2012, 47, 411–420. [Google Scholar] [CrossRef]
  48. Wang, H.; Ji, C.; Yang, J.; Wang, S.; Ge, Y. Towards a comprehensive optimization of the intake characteristics for side ported Wankel rotary engines by coupling machine learning with genetic algorithm. Energy 2022, 261, 125334. [Google Scholar] [CrossRef]
  49. Shi, C.; Chai, S.; Di, L.; Ji, C.; Ge, Y.; Wang, H. Combined experimental-numerical analysis of hydrogen as a combustion enhancer applied to wankel engine. Energy 2023, 263, 125896. [Google Scholar] [CrossRef]
  50. Wang, H.; Gan, H.; Theotokatos, G. Parametric investigation of pre-injection on the combustion, knocking and emissions behaviour of a large marine four-stroke dual-fuel engine. Fuel 2020, 281, 118744. [Google Scholar] [CrossRef]
  51. Richards, K.J.; Senecal, P.K.; Pomraning, E. CONVERGE 3.0 Manual; Convergent Science: Madison, WI, USA, 2020. [Google Scholar]
  52. Rahimi, A.; Fatehifar, E.; Saray, R.K. Development of an optimized chemical kinetic mechanism for homogeneous charge compression ignition combustion of a fuel blend of n-heptane and natural gas using a genetic algorithm. Proc. Inst. Mech. Eng. Part D J. Automob. Eng. 2010, 224, 1141–1159. [Google Scholar] [CrossRef]
  53. Liu, J.; Wang, J.; Zhao, H. Optimization of the injection parameters and combustion chamber geometries of a diesel/natural gas RCCI engine. Energy 2018, 164, 837–852. [Google Scholar] [CrossRef]
  54. Huang, H.; Lv, D.; Zhu, J.; Zhu, Z.; Chen, Y.; Pan, Y.; Pan, M. Development of a new reduced diesel/natural gas mechanism for dual-fuel engine combustion and emission prediction. Fuel 2019, 236, 30–42. [Google Scholar] [CrossRef]
  55. Wang, H.; Ji, C.; Shi, C.; Wang, S.; Yang, J.; Ge, Y. Investigation of the gas injection rate shape on combustion, knock and emissions behavior of a rotary engine with hydrogen direct-injection enrichment. Int. J. Hydrog. Energy 2021, 46, 14790–14804. [Google Scholar] [CrossRef]
  56. Bao, J.; Qu, P.; Wang, H.; Zhou, C.; Zhang, L.; Shi, C. Implementation of various bowl designs in an HPDI natural gas engine focused on performance and pollutant emissions. Chemosphere 2022, 303, 135275. [Google Scholar] [CrossRef]
  57. Huang, H.; Zhu, Z.; Chen, Y.; Chen, Y.; Lv, D.; Zhu, J.; Ouyang, T. Experimental and numerical study of multiple injection effects on combustion and emission characteristics of natural gas–diesel dual-fuel engine. Energy Convers. Manag. 2019, 183, 84–96. [Google Scholar] [CrossRef]
  58. Chen, Z.; Wang, L.; Wang, X.; Chen, H.; Geng, L.; Gao, N. Experimental study on the effect of water port injection on the combustion and emission characteristics of diesel/methane dual-fuel engines. Fuel 2022, 312, 122950. [Google Scholar] [CrossRef]
Figure 1. Schematic sector domain and computational mesh.
Figure 1. Schematic sector domain and computational mesh.
Sustainability 15 03610 g001
Figure 2. Comparison of predicted pressure profile under different meshes.
Figure 2. Comparison of predicted pressure profile under different meshes.
Sustainability 15 03610 g002
Figure 3. Model validation of in-cylinder pressure at various speeds.
Figure 3. Model validation of in-cylinder pressure at various speeds.
Sustainability 15 03610 g003
Figure 4. Comparison of in-cylinder pressure at hydrogen additional enrichment.
Figure 4. Comparison of in-cylinder pressure at hydrogen additional enrichment.
Sustainability 15 03610 g004
Figure 5. Comparison of combustion phases at hydrogen additional enrichment.
Figure 5. Comparison of combustion phases at hydrogen additional enrichment.
Sustainability 15 03610 g005
Figure 6. Comparison of emissions at hydrogen additional enrichment.
Figure 6. Comparison of emissions at hydrogen additional enrichment.
Sustainability 15 03610 g006
Figure 7. Comparison of in-cylinder pressure at hydrogen substitution enrichment.
Figure 7. Comparison of in-cylinder pressure at hydrogen substitution enrichment.
Sustainability 15 03610 g007
Figure 8. Comparison of HRR and combustion phases at hydrogen substitution enrichment.
Figure 8. Comparison of HRR and combustion phases at hydrogen substitution enrichment.
Sustainability 15 03610 g008
Figure 9. Comparison of emissions at hydrogen substitution enrichment.
Figure 9. Comparison of emissions at hydrogen substitution enrichment.
Sustainability 15 03610 g009
Figure 10. Comparison of in-cylinder pressure with various pilot injection timings.
Figure 10. Comparison of in-cylinder pressure with various pilot injection timings.
Sustainability 15 03610 g010
Figure 11. Comparison of HRR and combustion phases with various pilot injection timings.
Figure 11. Comparison of HRR and combustion phases with various pilot injection timings.
Sustainability 15 03610 g011
Figure 12. Comparison of emissions with various pilot injection timings.
Figure 12. Comparison of emissions with various pilot injection timings.
Sustainability 15 03610 g012
Table 1. Mathematical models adopted in this research.
Table 1. Mathematical models adopted in this research.
RegionType
TurbulenceRNG k ε model
Wall heat transferO’Rourke and Amsden model
Spray breakupKH-RT model
EvaporationFrossling model
Droplet collisionO’Rourke’s model
CombustionSAGE
NOx formationExtended Zel’dovich mechanism
Soot formationHiroyasu model
Table 2. Summary of different calculated schemes.
Table 2. Summary of different calculated schemes.
Case GroupHydrogen Energy Ratio/%Relative Injection Timing/°CA
Hydrogen enrichment0% 10
10% 20
20% 20
Hydrogen substitution30%0
60%0
90%0
Relative injection timing0%0
90%0
90%5
90%10
90%15
1 pure diesel, 2  H E R .
Table 3. Comparison of PFP, corresponding CA position, and IMEP at hydrogen additional enrichment.
Table 3. Comparison of PFP, corresponding CA position, and IMEP at hydrogen additional enrichment.
ParameterPFP (bar)PFP_CA (°CA, aTDC)IMEP (bar)
D10021.757.2621.53
D100 + H1022.497.4323.40
D100 + H2023.517.7325.33
Table 4. Contour of temperature distribution at hydrogen additional enrichment.
Table 4. Contour of temperature distribution at hydrogen additional enrichment.
CADD100 (K)D100 + H10 (K)D100 + H20 (K)
−4 °CA aTDCSustainability 15 03610 i001Sustainability 15 03610 i002Sustainability 15 03610 i003
TDCSustainability 15 03610 i004Sustainability 15 03610 i005Sustainability 15 03610 i006
4 °CA aTDCSustainability 15 03610 i007Sustainability 15 03610 i008Sustainability 15 03610 i009
8 °CA aTDCSustainability 15 03610 i010Sustainability 15 03610 i011Sustainability 15 03610 i012
16 °CA aTDCSustainability 15 03610 i013Sustainability 15 03610 i014Sustainability 15 03610 i015
24 °CA aTDCSustainability 15 03610 i016Sustainability 15 03610 i017Sustainability 15 03610 i018
Table 5. Comparison of PFP, corresponding CA position, and IMEP at hydrogen additional enrichment.
Table 5. Comparison of PFP, corresponding CA position, and IMEP at hydrogen additional enrichment.
ParameterPFP (bar)PFP_CA (°CA, aTDC)IMEP (bar)
D100-H021.757.2621.53
D70-H3023.935.9222.00
D40-H6028.331.8522.01
D10-H9030.13−0.4821.76
Table 6. Contour of temperature distribution at hydrogen substitution enrichment.
Table 6. Contour of temperature distribution at hydrogen substitution enrichment.
CADD100-H0 (K)D70-H30 (K)D40-H60 (K)D10-H90 (K)
−4 °CA aTDCSustainability 15 03610 i019Sustainability 15 03610 i020Sustainability 15 03610 i021Sustainability 15 03610 i022
TDCSustainability 15 03610 i023Sustainability 15 03610 i024Sustainability 15 03610 i025Sustainability 15 03610 i026
4 °CA aTDCSustainability 15 03610 i027Sustainability 15 03610 i028Sustainability 15 03610 i029Sustainability 15 03610 i030
12 °CA aTDCSustainability 15 03610 i031Sustainability 15 03610 i032Sustainability 15 03610 i033Sustainability 15 03610 i034
24 °CA aTDCSustainability 15 03610 i035Sustainability 15 03610 i036Sustainability 15 03610 i037Sustainability 15 03610 i038
Table 7. Comparison of PFP, corresponding CA position, and IMEP with various pilot injection timings.
Table 7. Comparison of PFP, corresponding CA position, and IMEP with various pilot injection timings.
ParameterPFP (bar)PFP_CA (°CA, aTDC)IMEP (bar)
H0-I021.757.2621.53
H90-I030.13−0.4821.76
H90-I529.572.5222.00
H90-I1025.559.3222.27
H90-I1515.3717.0321.20
Table 8. Contour of temperature distribution with various pilot injection timings.
Table 8. Contour of temperature distribution with various pilot injection timings.
CADH90-I0 (K)H90-I5 (K)H90-I10 (K)H90-I15 (K)
−4 °CA aTDCSustainability 15 03610 i039Sustainability 15 03610 i040Sustainability 15 03610 i041Sustainability 15 03610 i042
TDCSustainability 15 03610 i043Sustainability 15 03610 i044Sustainability 15 03610 i045Sustainability 15 03610 i046
8 °CA aTDCSustainability 15 03610 i047Sustainability 15 03610 i048Sustainability 15 03610 i049Sustainability 15 03610 i050
16 °CA aTDCSustainability 15 03610 i051Sustainability 15 03610 i052Sustainability 15 03610 i053Sustainability 15 03610 i054
24 °CA aTDCSustainability 15 03610 i055Sustainability 15 03610 i056Sustainability 15 03610 i057Sustainability 15 03610 i058
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, B.; Wang, H.; Wang, S. Computational Investigation of Combustion, Performance, and Emissions of a Diesel-Hydrogen Dual-Fuel Engine. Sustainability 2023, 15, 3610. https://doi.org/10.3390/su15043610

AMA Style

Zhang B, Wang H, Wang S. Computational Investigation of Combustion, Performance, and Emissions of a Diesel-Hydrogen Dual-Fuel Engine. Sustainability. 2023; 15(4):3610. https://doi.org/10.3390/su15043610

Chicago/Turabian Style

Zhang, Bo, Huaiyu Wang, and Shuofeng Wang. 2023. "Computational Investigation of Combustion, Performance, and Emissions of a Diesel-Hydrogen Dual-Fuel Engine" Sustainability 15, no. 4: 3610. https://doi.org/10.3390/su15043610

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