Next Article in Journal
Assessment of Heavy Metal Contamination and Ecological Risk in Soil within the Zheng–Bian–Luo Urban Agglomeration
Next Article in Special Issue
Study on the Evolution of Physicochemical Properties of Carbon Black at Different Regeneration Stages of Diesel Particulate Filters Regenerated by Non-Thermal Plasma
Previous Article in Journal
Plasma Processing of Rubber Powder from End-of-Life Tires: Numerical Analysis and Experiment
Previous Article in Special Issue
Using a One-Dimensional Convolutional Neural Network with Taguchi Parametric Optimization for a Permanent-Magnet Synchronous Motor Fault-Diagnosis System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Constructing a Skeletal Iso-Propanol–Butanol–Ethanol (IBE)–Diesel Mechanism Using the Decoupling Method

1
School of Mechanical and Energy Engineering, Shaoyang University, Shaoyang 422000, China
2
Key Laboratory of Hunan Province for Efficient Power System and Intelligent Manufacturing, Shaoyang University, Shaoyang 422000, China
3
CSSC Marine Power Co., Ltd., Shanghai 201208, China
*
Authors to whom correspondence should be addressed.
Processes 2024, 12(5), 995; https://doi.org/10.3390/pr12050995
Submission received: 5 April 2024 / Revised: 25 April 2024 / Accepted: 7 May 2024 / Published: 14 May 2024
(This article belongs to the Special Issue Clean Combustion and Emission in Vehicle Power System, 2nd Edition)

Abstract

:
In recent years, biofuels have gained considerable prominence in response to growing concerns about resource scarcity and environmental pollution. Previous investigations have revealed that the appropriate blending of iso-propanol–butanol–ethanol (IBE) into diesel significantly improves both the c combustion efficiency and emission performance of internal combustion engines (ICEs). However, the combustion mechanism of IBE–diesel for the numerical studies of engines has not reached maturity. In this study, a skeletal IBE–diesel multi-component mechanism, comprising 157 species and 603 reactions, was constructed using the decoupling method. It was formulated by amalgamating the reduced fuel-related sub-mechanisms derived from diesel surrogates (n-dodecane, iso-cetane, iso-octane, toluene, and decalin) and n-butanol, along with the detailed core sub-mechanisms of C1, C2, C3, CO, and H2. The constructed mechanism is capable of better matching the physical and chemical properties of actual diesel fuel. Extensive validation, including ignition delay, laminar flame speed, a premixed flame species profile, and engine experimental data, confirms the reliability of the mechanism in engine numerical studies. Subsequent investigations reveal that as the IBE blend ratio and EGR rate increase, the ignition delay exhibits an increase, while the combustion duration experiences a decrease. Blending IBE into diesel, along with a specific EGR rate, proves effective in simultaneously reducing NOx and soot emissions.

1. Introduction

Recently, both Europe and China have introduced tougher emission regulations to reduce the environmental impact of the transport sector. Contemporary diesel engines require increased energy efficiency and decreased toxic emissions to help address energy security issues and meet stringent emission laws [1,2,3]. Investigations into alternative fuels for internal combustion engines (ICEs) have become significant lately due to growing concerns about ecological issues. Extensive research has been conducted on the utilization of bio-butanol as a fuel or blending component in ICEs [4,5,6,7]. However, the extraction and fermentation process of bio-butanol reduces its economic viability. During the fermentation process, the byproduct acetone–butanol–ethanol (ABE) has attracted scholarly attention. Research indicates that ABE is a viable alternative fuel suitable for use in ICEs. However, acetone is highly corrosive, with a very low flash point [8], making it susceptible to flash combustion and challenging to transport and store. Currently, there exists a mature transformation technology that utilizes strains with gene editing to convert ABE into iso-propanol–butanol–ethanol (IBE) [9,10,11]. Moreover, the cost of this transformation technology is significantly lower than that of separating and purifying biological butanol. The calorific value, viscosity, and flash point of iso-propanol are higher than those of acetone, with low corrosivity, thereby compensating for the shortcomings associated with acetone. Therefore, the utilization of iso-propanol as an alternative fuel to diesel in ICEs holds promising prospects for development.
The combustion and emission characteristics of IBE in ICEs have been explored by many scholars to demonstrate its generalizability and feasibility as a biofuel. Intensive investigations into the combustion and emission characteristics of IBE and aqueous IBE blended with gasoline as fuels were conducted by Li et al. [12,13] on a spark ignition engine. The results indicate that the addition of IBE can help improve harmful emissions such as nitrogen oxide (NOx) and carbon monoxide (CO) while simultaneously enhancing fuel efficiency. Furthermore, the combustion and emission characteristics of IBE blended with diesel fuel were evaluated by Lee et al. [14,15], who analyzed parameters such as the ignition delay and heat release rate at various blend ratios. The results indicated that, under certain conditions, the addition of IBE increased the ignition delay, while the natural flame luminosity of the blends decreased with the increase in the proportion of IBE. This suggests that the addition of IBE has the potential to reduce soot emissions. However, conducting an in-depth experimental study will certainly consume a lot of effort and resources, so performing numerical simulations of IBE combustion as a fuel in ICEs becomes a suitable option. Hu et al. [16] and Li et al. [17] simplified the chemical reaction kinetics of IBE–diesel by employing a mixture of n-heptane and toluene as diesel, applying various reduction methods to reduce the detailed mechanism. After extensive validation, the mechanism constructed by them can match well with the experimental ignition delay and laminar flame speed. However, diesel generally consists of hundreds or even an abundance of distinct components [18]. A substitute containing cycloalkanes, alkanes, and aromatics can replicate the oxidation behaviors observed in diesel [19]. The mechanisms of benchmark fuel fail to analyze the oxidation of these four types, resulting in uniformly fluctuating heat release behaviors [20]. Specifically, due to the lack of aromatic compounds, benchmark fuel produces less soot compared to diesel [21]. Recently, Yu et al. [22] created a model using the following components: decalin (C10H18), iso-cetane (IC16H34), n-dodecane (NC12H26), toluene (C7H8), and iso-octane (IC8H18). This substitute can simultaneously satisfy the physicochemical properties of real diesel fuel.
In this study, a novel skeletal model was constructed with the aim of bridging the gap between the existing IBE–diesel combustion chemical reaction mechanisms. The model is characterized by multi-element substitutions, which are consistent with the physicochemical behavior of actual diesel fuel. The model underwent comprehensive validation, including the evaluation of ignition delay, laminar flame velocity, premixed flame species distribution, and numerical engine studies.

2. Mechanism Construction

2.1. Strategy

The decoupling method divides the mechanism into three parts: The detailed H2/CO/C1 sub-mechanism is used for predicting the flame propagation rate, heat release rate, and concentration of the major components. The skeletal C4–Cn sub-mechanism is employed for predicting fuel consumption and the stagnation period. The simplified C2–C3 sub-mechanism serves as a transition mechanism between the above two sub-mechanisms. By employing this method, the scale of the mechanism can be reduced, thereby enhancing the efficiency of numerical simulation while ensuring the reliability of the mechanism [23]. As illustrated in Figure 1, the construction of a skeletal mechanism for simulating the combustion of IBE–diesel blends is presented in this section. NC12H26, IC16H34, and IC8H18 represent long-chain hydrocarbons, while C7H8 and C10H18 represent aromatic hydrocarbons. These five components were utilized to characterize diesel, and the physicochemical characteristics of the mixture nearly approximate those of actual diesel [22]. Firstly, the top layer sub-mechanism for iso-propanol, n-butanol, ethanol, and diesel was constructed individually. This was then combined with the PAH model and NOx model, and ultimately coupled with the C1, C2, C3, CO, and H2 sub-mechanism, resulting in the establishment of a comprehensive multi-element skeletal mechanism for IBE–diesel blends.

2.2. Sub-Mechanism of Diesel

2.2.1. Sub-Mechanism of n-Dodecane, Iso-Octane, and Iso-Cetane

NC12H26, IC16H34, and IC8H18 are classified as long-chain alkane compounds. In this study, the reaction rate constants of specific low-temperature reactions within the iso-octane sub-mechanism, encompassing dehydrogenation, iso-octyl radical oxidation, isomerization, further oxidation subsequent to isomerization, and high-temperature iso-octane and iso-octyl cracking, relied on the mechanism proposed by Dong et al. [24], while the mechanisms governing OH release and ketone hydroperoxide decomposition were primarily derived from Chang et al. [23]. The detailed mechanism of n-dodecane was derived from Ref. Dagaut et al. [25], and the fuel-related mechanism of n-dodecane was developed with reference to it. The CRECK model [26] was employed in formulating the top-level mechanism for iso-cetane, while the high-temperature pyrolysis products and related reactions of iso-cetane were characterized using the reduced mechanism proposed by Fan et al. [27]. The overall reaction pathways of long-chain hydrocarbons exhibit similarity, as illustrated in Figure 2.

2.2.2. Sub-Mechanism of Decalin

The decalin sub-model pertains to the detailed model established by Dagaut et al. [28], which accounts for the reaction distinctions between low and high temperatures and accurately depicts the oxidation of decalin across a broad temperature range. The primary reaction pathway of decalin is depicted in Figure 3. Owing to the symmetrical structure of decalin, three free radicals can be produced during dehydrogenation, coalescing into RDECALIN. At low temperatures, RDECALIN primarily undergoes oxidation reactions, and the alkyl peroxy radical generated through oxidation undergoes isomerization with RDECOO, resulting in the formation of the hydroperoxyalkyl QDECOOH. QDECOOH produces the C10H17O4 isomer via oxidation, which is collectively represented as ZDECA. ZDECA eliminates hydroxyl and dissociates into KHDECA. Ultimately, the low-temperature reaction chain concludes through the decomposition reaction of KHDECA. At high temperatures, RDECALIN predominantly yields C5H8 and the cyclopentyl cyclopentene DCYC5. C5H8 undergoes dehydrogenation, resulting in the formation of the cyclopentadiene CYC5H6. CYC5H6 reacts with DCYC5, producing C5H5 and A1, both crucial precursors for PAH formation.

2.2.3. Sub-Mechanism of Toluene

The fuel-related sub-mechanism for toluene is derived from the skeletal diesel mechanism proposed by Dong et al. [24]. Toluene oxidation involves sequential steps, including dehydrogenation to generate benzyl and the removal of methyl groups to produce A1. Benzyl oxidation results in the formation of C6H5CHO, which undergoes further dehydrogenation to produce phenyl. Concurrently, A1 undergoes dehydrogenation to form phenyl. Phenyl is oxidized to generate C6H5O, resulting in ring opening and the production of a small cyclic hydrocarbon, C5H5, which decomposes into C0–C3 substances or smaller species after oxidation (Figure 4).

2.3. Sub-Mechanism of IBE

The sub-mechanism of iso-propanol primarily pertains to the mechanism conducted by Man et al. [29], which encompasses 238 components and 1448 reactions, providing detailed insights into the reaction pathways of the two propanol isomers. The primary products derived from iso-propanol include acetone and propylene. As shown in Figure 5, in the decomposition reaction process, TC3H6OH and IC3H6OH emerge as significant products resulting from the dehydrogenation of iso-propanol. Among these, acetone is generated through the dehydrogenation of TC3H6OH, followed by subsequent decomposition. IC3H6OH eliminates hydroxyl to yield propylene. Therefore, both these products and their associated reactions are preserved.
Gaurav et al. [30] developed an ethanol mechanism comprising 113 components and 710 reactions. This mechanism was utilized as a reference to formulate the fuel-related mechanism for ethanol. The dehydrogenation product SC2H4OH from ethanol is also derived by eliminating a methyl group from iso-propanol (see Figure 5). Additional decomposition and isomerization reactions associated with SC2H4OH are incorporated into the ethanol model.
The simplified model for n-butanol was derived from the one comprising four butanol isomers proposed by Chang et al. [31]. It is observed from Figure 5 that n-butanol undergoes dehydrogenation, resulting in the production of n-butyl radicals at low temperatures. Subsequently, oxidation and isomerization processes generate NC4H7OHOO2. The ketohydroperoxide C4ket is produced through the release of OH from NC4H7OHO2. C4ket then decomposes into smaller molecules, including CH2CO, CH2OH, CH2O, and OH. At elevated temperatures, n-butanol undergoes direct decomposition into ethylene and several C0–C3 species. Simultaneously, the β-break reaction of C4H8OH radicals yields ethyl and vinyl alcohol radicals [32]. The highly unstable ethanol undergoes rapid isomerization to form acetaldehyde [33], which further decomposes into smaller species. The ethyl radical undergoes dehydrogenation, leading to the generation of ethylene. Chang et al. [31] consolidated the cleavage reaction at high temperatures into a single reaction, where the primary pyrolysis byproducts include ethylene and OH radicals.

2.4. Soot and NOx Models

The model depicting the generation of polycyclic aromatic hydrocarbons (PAHs), as a soot precursor, holds significant importance in predicting the behavior of soot emissions. The mechanism comprises a sequence of fundamental reaction steps starting from acetylene and hydrogen, progressing from the initial stages of ring formation to the development of the aromatic ring and, consequently, the phenyl group. The process involves free radical formation, followed by successive hydrogen absorption and acetylene addition (HACA mechanism) stages, leading to the generation of the aromatic chain. These details are incorporated into the mechanism description for the formation of higher aromatic hydrocarbons. The pathways leading to higher rings follow a similar pattern, with the distinction being an increase in the number of rings in each molecule. This is illustrated as follows, where soot is represented by C(s) [34]. Soot ultimately originates from the long-chain acetylene C6H2 and heptadiene A2R5. To accurately reflect the pattern of soot emissions in engine combustion simulations, the soot mechanism presented by Vishwanathan and Reitz [35] was adapted and integrated into the existing model to elucidate the oxidation process of soot.
The reduced model constructed by Bergman et al. [36] was chosen for the NOx formation model. This model encompasses 13 reactions, encompassing the formation of NO, the low-temperature combustion chain resulting in the production of N2O and subsequently NO, and the further oxidation of NO to NO2. Extensive testing has validated its capability to precisely forecast NOx emissions in engine scenarios.

3. Mechanism Validation

3.1. Ignition Delay

Ignition delay is characterized as the duration required for a temperature rise of 400 K compared to the starting temperature. In Figure 6, a comparison is made between the empirically measured and computationally determined ignition delay for n-dodecane [37,38], iso-octane [39,40], and iso-cetane [41]. The accuracy of the prediction results largely conforms to the specified requirements. In Figure 6a,b, it is evident that the current model adeptly captures the negative temperature coefficient (NTC) characteristics of n-dodecane; the predicted n-dodecane ignition delays at various pressures were in agreement with the experimental measurements. In Figure 6c,d, the ignition delay of iso-octane is presented under various pressure and equivalence ratio conditions. At T = 1000~1250 K, φ = 1.0 and P = 40 atm and T = 850~1000 K, φ = 1.5, and P = 40 atm, the results show that the simulated ignition delay of iso-octane is marginally lower than the measured data. However, as can be seen from Figure 6e,f, the simulated value for iso-cetane is slightly higher than the measured value. This is due to the fact that during the fuel-related mechanism development for iso-cetane, a crucial decomposition reaction, specifically IC16KET => IC8H17 + 3C2H4 + CH2O + OH + CO, involving the ketide of iso-cetane was considered.
Figure 7 illustrates the simulated and experimental high-temperature ignition delays of toluene at φ = 0.25, 0.5, and 1.0 and 4 atm [42]. Measuring the ignition delay of toluene at low temperatures poses challenges; hence, this study primarily focuses on comparing the high-temperature ignition delay. Despite some errors in the simulation results, they exhibit a good fit with the experimental data and meet the required prediction accuracy.
Figure 8 illustrates the comparison between the measured [43] and simulated ignition delays of decalin across 770–1250 K, encompassing the NTC region. The comparisons are made at 20 and 50 atm, as well as φ = 0.5 and 1.0. It is evident that, at 50 atm and φ = 1.0, a noticeable NTC effect occurs. This phenomenon arises because, in the NTC region, the reaction rate of O2 addition to form an alkyl group is significantly influenced by pressure. The mechanism constructed in this study primarily focuses on small molecules related to chain hydrocarbons and their reactions. It does not account for the primary small molecular substances and reactions that occur after the decomposition of the bicyclic structure of decalin. As a result, there is a deviation between the predicted results of the constructed mechanism and the measured data in the NTC region.
After validating the single-component model, the ignition delay of diesel was also confirmed. As depicted in Figure 9a, the simulation results for US#2 diesel align well with the measured data. However, it is essential to note that the experimental results primarily concentrated on the high-temperature region, and no experimental data were provided for the low-temperature region. Gowdagiri et al. [44] conducted measurements of the hysteretic ignition period for F-76 diesel in a shock tube. It is found from Figure 9b that the simulation exhibited good performance at both pressures, with only minor errors observed in the NTC region at P = 20 atm. One possible reason is the complexity of the actual composition of diesel. Despite using various categories of components to characterize it and considering several diesel fuel properties, it cannot entirely replicate real diesel fuel, leading to inevitable deviations. Additionally, during the mechanism construction, it is significantly reduced to streamline its size and facilitate 3D calculations, introducing some errors in the process. However, in general, the simulated ignition delay of diesel has met the anticipated level of accuracy.
The ignition delay for iso-propanol combustion in an oxygen–argon mixture has been obtained from the research conducted by Man et al. [29]. Figure 10 illustrates a comparison between the experimental ignition delay under various pressures and equivalence ratios and the simulated ignition delay by the mechanism. At 1.2 atm, the calculated data slightly exceed the experimental values, though the overall fitting result is satisfactory.
The experimental ignition delay data for ethanol/N2/O2/Ar mixtures under various conditions have been sourced from Mittal et al. [30] and Dunphy and Simmie [45]. It can be seen from Figure 11a that the current mechanism consistently aligns with experimental values at both high and low temperatures, with any slight deviations falling within an acceptable range. The experimental data for n-butanol are derived from the measurements conducted by Heufer et al. [46] and Vrankx et al. [47]. As depicted in Figure 11b, the overall fitting result is satisfactory; however, at lower temperatures, particularly when P = 20 bar, a minor discrepancy is observed. This discrepancy may be attributed to the construction of the mechanism, wherein five isomers of n-butyl are generated after n-butanol dehydrogenation, and only the most crucial one is retained to minimize the mechanism size, thereby decreasing n-butanol consumption paths and delaying ignition at low temperatures. Generally, the prediction accuracy for n-butanol remains acceptable.

3.2. Laminar Flame Speed

Figure 12 presents the experimental laminar flame speed (LFS) of five diesel components: n-dodecane [48,49], iso-cetane [27], iso-octane [48,50], toluene [51], and decalin [52], alongside the corresponding predicted outcomes. Clearly, the simulated laminar flame speed meets the requirements for n-dodecane, iso-octane, and toluene. A minor disparity persists between the simulated and experimental data for decalin, particularly noticeable at φ = 0.8–1.1, as illustrated in Figure 12c. This arises due to the unique bicyclic structure of decalin, differing from alkanes, as indicated by the current mechanism that inadequately describes the reactions of small species (C0–C3) in relation to it. Nevertheless, there has been a significant improvement in the simulated laminar flame speed of decalin in comparison to the previous model developed by Nakamura et al. [53]. It remains within an acceptable margin of error. The measurement of the laminar flame speed of iso-cetane is seldom encountered in the existing literature. The measured results obtained from a modified CFR octane rating engine, along with the mechanism developed by Dooley et al. [54], were employed to confirm the validity of the model. Figure 12e exhibits some deviations, primarily stemming from the technique’s original design for species identification rather than quantification. In summary, the noted discrepancies notwithstanding, the overall agreement instills confidence in the capacity of the proposed model to effectively depict the oxidation characteristics of the kinetics-controlling pathways for small-molecular-weight compounds in diesel.
As depicted in Figure 13a, the laminar flame speed of iso-propanol has been confirmed across a pressure range of 1–10 atm at 423 K [55]. Figure 13b illustrates a good fit of ethanol between the experimental and simulated laminar flame speed at 1 bar and various initial temperatures [56,57,58,59]. In Figure 13c, a comparison is presented between the simulated laminar flame speed of n-butanol and the measured data [60] at T = 423 K under varying pressure conditions. The simulated value aligns well with the measured one. The verification results indicate a close correspondence between the calculated laminar flame speed of the IBE sub-mechanism and the experimental measurement value. This alignment underscores the effective coupling of the top mechanism of IBE, as constructed in this paper, with the bottom mechanism and transition mechanism.

3.3. Premixed Flame Species Profile

In Figure 14, the mole fraction of primary products from the premixed flame of iso-octane in the O2 and AR mixture at 1 bar and φ = 1.47 is illustrated. It can be seen that the predictions of the mechanism for the molar fractions of most species are very close to the experimental measurements. The figure underscores the reliability of the current iso-octane sub-mechanism [61].
Figure 15 presents the mass distribution concentration of the n-dodecane premixed flame within 550–1150 K at 10 bar and various equivalence ratios [62]. The consumption of n-dodecane and the concentration changes of key products align closely with the experimental values under two equivalence ratios.
Dagaut and Hadj Ali [63] investigated the concentration of reactants and end products in the oxidation of iso-cetane in a JSR. The experiment was conducted at 10 atm, φ = 0.5–1.0, and 770–1070 K. In Figure 16, a comparison is drawn between the simulated and measured concentrations of iso-cetane, CO2, CO, and H2O. The figure illustrates the accurate reproduction of iso-cetane consumption under all operating conditions. Notably, at low temperatures, the predicted value of CO exceeds the experimental value, possibly due to the absence of initial fuel decomposition reactions in the current mechanism. Despite this, the overall trend is well captured.
Figure 17a shows a comparison between the calculated values and experimental results [64] of the concentration change in major substances in a toluene premixed flame. The mechanism accurately captures the concentration distribution of various components of toluene. In Figure 17b, the calculated premixed flame concentration of decalin is shown. While the overall trend aligns well with the experimental results [65], there is a slight deviation. Possible reasons include the simplification of the mechanism size through the removal of more components and related reactions during reduction. Additionally, the aggregation of reactions of isomers into one type, for the sake of simplification, might introduce some errors. Another factor to consider is the potential fluctuation of temperatures near the burner surface during experiments, influenced by radiant heat loss and cooling [66].
Mati et al. [67] conducted a study on diesel oxidation under conditions of 10 atmospheres and in a JSR with a fixed residence time of 0.5 s and an equivalence ratio of 0.5. The test fuel used was a synthetic diesel oil mixed with various oil solvents, resembling commercial diesel. In Figure 18, it is evident that the existing DFS mechanism impeccably reproduces the O2 concentration history. Similarly, the current mechanism accurately predicts the peak value of CO history, with a slight overestimation observed at low temperatures compared to the experimental values.
The experiment involving the premixed flame mass concentration distribution of iso-propanol was conducted in a JSR [68]. In Figure 19a, a comparison is made between the concentration distribution of iso-propanol, CO, CO2, and H2O in the experiment and simulation. Figure 19b illustrates the component concentration distribution of ethanol during combustion in an O2 and AR mixture. The predicted values generally align well with the experimental results. The component concentration distribution test results for n-butanol are sourced from Hansen et al. [69]. It is evident from Figure 19 that the IBE sub-mechanism yields accurate predictions, effectively capturing the concentration changes of small molecular components during combustion.

3.4. Validation of Engine Combustion

To enhance the reliability of the current mechanism in three-dimensional simulations, a comparative study was conducted against the experiments by Li et al. [15]. Three fuel blends were examined: pure diesel and IBE15 (defined as 85 vol.% diesel–15 vol.% IBE mixed fuel, with IBE component ratios I:B:E = 3:6:1). The injection pressure was maintained at 100 MPa. Tests were performed on a single-cylinder common rail direct injection diesel engine (AVL 5402). The simulation results for the in-cylinder pressure and heat release rate were then compared with the corresponding experimental data. Figure 20 shows the established calculation grid for the fifth engine combustion chamber at the top dead center (TDC).
The accuracy of the three-dimensional model, coupled with the current mechanism, was verified by comparing the cylinder pressure and heat release rate between the experimental and simulated data. Table 1 presents the engine specifications and operation conditions. As illustrated in Figure 21, the simulation results demonstrate a generally consistent trend with the measured values in terms of the in-cylinder pressure and heat release rate. Any slight discrepancies could be attributed to the omission of turbulence effects on chemical kinetics. The figure highlights that the mechanism developed in this study, when coupled with the numerical model, effectively reproduces the combustion characteristics of IBE–diesel blends in the engine.

4. Conclusions

The current research introduces a realistic skeletal mechanism for diesel–IBE combustion within an engine, employing the decoupling methodology. The hierarchical construction of the mechanism involved integrating the fuel-specific sub-model of IBE with diesel surrogates such as n-dodecane, iso-cetane, iso-octane, toluene, and decalin, alongside the core sub-mechanisms of C1, C2, C3, CO, and H2, generating a comprehensive system including 157 species and 603 reactions. Compared with previous studies, the mechanism constructed in this study can better match the physical and chemical properties of actual diesel fuel, and it has higher reliability and generalizability. The proposed mechanism underwent thorough validation, encompassing ignition delay, laminar flame speed, and a premixed flame species profile, as well as the in-cylinder pressure and heat release rate within the engine. The overall congruence observed between the predicted and measured data attests to the viability and applicability of the mechanism in real-world engine applications. In the future, it will be necessary to validate the accuracy of the mechanism in predicting important emissions such as NOx and soot based on experimental data.

Author Contributions

Conceptualization, Y.M., S.Z. and J.Z.; methodology, Y.M. and J.Z.; software, Y.M. and S.Z.; validation, Y.M., S.Z. and J.Z.; formal analysis, Y.M.; investigation, Y.M.; resources, Y.M. and J.Z.; data curation, Y.M.; writing—original draft preparation, Y.M.; writing—review and editing, Y.M., J.Z., J.F. and W.Y.; supervision, W.Y.; project administration, W.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (Grant No. 52076141), funder: W.Y., Hunan Provincial Natural Science Foundation of China (Grant No. 2023JJ50262), funder: Y.M., and Postgraduate Scientific Research Innovation Project of Hunan Province, (Grant No. CX20231294), funder: S.Z.

Data Availability Statement

The data used to support the findings of this study are available from the corresponding author upon request.

Conflicts of Interest

The author Junhong Zhao is employed by the company CSSC Marine Power Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest.

References

  1. Meng, Z.; Chen, C.; Li, J.; Fang, J.; Tan, J.; Qin, Y.; Jiang, Y.; Qin, Z.; Bai, W.; Liang, K. Particle Emission Characteristics of DPF Regeneration from DPF Regeneration Bench and Diesel Engine Bench Measurements. Fuel 2020, 262, 116589. [Google Scholar] [CrossRef]
  2. Tan, Y.; Kou, C.; Jiaqiang, E.; Feng, C.; Han, D. Effect of Different Exhaust Parameters on Conversion Efficiency Enhancement of a Pd–Rh Three-Way Catalytic Converter for Heavy-Duty Natural Gas Engines. Energy 2024, 292, 130483. [Google Scholar] [CrossRef]
  3. Han, D.; Jiaqiang, E.; Feng, C.; Han, C.; Kou, C.; Tan, Y.; Peng, Y.; Wei, L. Experimental and Simulation Investigation on the Different Iron Content Beta Zeolite for Controlling the Cold-Start Hydrocarbon Emission from a Gasoline Vehicle. Energy 2024, 294, 130954. [Google Scholar] [CrossRef]
  4. Szwaja, S.; Naber, J.D. Combustion of n-Butanol in a Spark-Ignition IC Engine. Fuel 2010, 89, 1573–1582. [Google Scholar] [CrossRef]
  5. Lujaji, F.; Bereczky, A.; Janosi, L.; Novak, C.; Mbarawa, M. Cetane Number and Thermal Properties of Vegetable Oil, Biodiesel, 1-Butanol and Diesel Blends. J. Therm. Anal. Calorim. 2010, 102, 1175–1181. [Google Scholar] [CrossRef]
  6. Rakopoulos, D.C.; Rakopoulos, C.D.; Giakoumis, E.G.; Dimaratos, A.M.; Kyritsis, D.C. Effects of Butanol–Diesel Fuel Blends on the Performance and Emissions of a High-Speed DI Diesel Engine. Energy Convers. Manag. 2010, 51, 1989–1997. [Google Scholar] [CrossRef]
  7. Rakopoulos, D.C.; Rakopoulos, C.D.; Papagiannakis, R.G.; Kyritsis, D.C. Combustion Heat Release Analysis of Ethanol or n-Butanol Diesel Fuel Blends in Heavy-Duty DI Diesel Engine. Fuel 2011, 90, 1855–1867. [Google Scholar] [CrossRef]
  8. Tummala, S.B.; Welker, N.E.; Papoutsakis, E.T. Design of Antisense RNA Constructs for Downregulation of the Acetone Formation Pathway of Clostridium Acetobutylicum. J. Bacteriol. 2003, 185, 2973. [Google Scholar] [CrossRef]
  9. Dusséaux, S.; Croux, C.; Soucaille, P.; Meynial-Salles, I. Metabolic Engineering of Clostridium Acetobutylicum ATCC 824 for the High-Yield Production of a Biofuel Composed of an Isopropanol/Butanol/Ethanol Mixture. Metab. Eng. 2013, 18, 1–8. [Google Scholar] [CrossRef]
  10. Bankar, S.B.; Jurgens, G.; Survase, S.A.; Ojamo, H.; Granström, T. Enhanced Isopropanol–Butanol–Ethanol (IBE) Production in Immobilized Column Reactor Using Modified Clostridium Acetobutylicum DSM792. Fuel 2014, 136, 226–232. [Google Scholar] [CrossRef]
  11. Jang, Y.; Malaviya, A.; Lee, J.; Im, J.A.; Lee, S.Y.; Lee, J.; Eom, M.; Cho, J.; Seung, D.Y. Metabolic Engineering of Clostridium Acetobutylicum for the Enhanced Production of Isopropanol-butanol-ethanol Fuel Mixture. Biotechnol. Prog. 2013, 29, 1083–1088. [Google Scholar] [CrossRef]
  12. Li, Y.; Meng, L.; Nithyanandan, K.; Lee, T.H.; Lin, Y.; Lee, C.F.; Liao, S. Combustion, Performance and Emissions Characteristics of a Spark-Ignition Engine Fueled with Isopropanol-n-Butanol-Ethanol and Gasoline Blends. Fuel 2016, 184, 864–872. [Google Scholar] [CrossRef]
  13. Li, Y.; Chen, Y.; Wu, G.; Liu, J. Experimental Evaluation of Water-Containing Isopropanol-n-Butanol-Ethanol and Gasoline Blend as a Fuel Candidate in Spark-Ignition Engine. Appl. Energy 2018, 219, 42–52. [Google Scholar] [CrossRef]
  14. Li, G.; Liu, Z.; Lee, T.H.; Lee, C.F.; Zhang, C. Effects of Dilute Gas on Combustion and Emission Characteristics of a Common-Rail Diesel Engine Fueled with Isopropanol-Butanol-Ethanol and Diesel Blends. Energy Convers. Manag. 2018, 165, 373–381. [Google Scholar] [CrossRef]
  15. Lee, T.H.; Hansen, A.C.; Li, G.; Lee, T. Effects of Isopropanol-Butanol-Ethanol and Diesel Fuel Blends on Combustion Characteristics in a Constant Volume Chamber. Fuel 2019, 254, 115613. [Google Scholar] [CrossRef]
  16. Hu, J.; Abubakar, S.; Li, Y. A Novel Reduced I-Propanol-n-Butanol-Ethanol (IBE)/Diesel Mechanism for Engine Combustion and Emissions Prediction. Fuel 2020, 278, 118291. [Google Scholar] [CrossRef]
  17. Li, Y.; Lou, B.; Abubakar, S.; Wu, G. Skeletal Mechanism for I-Propanol-n-Butanol-Ethanol (IBE) and n-Butanol Combustion in Diesel Engine. Fuel 2021, 302, 121136. [Google Scholar] [CrossRef]
  18. Farrell, J.T.; Cernansky, N.P.; Dryer, F.L.; Law, C.K.; Friend, D.G.; Hergart, C.A.; McDavid, R.M.; Patel, A.K.; Mueller, C.J.; Pitsch, H. Development of an Experimental Database and Kinetic Models for Surrogate Diesel Fuels; SAE International: Warrendale, PA, USA, 2007. [Google Scholar]
  19. Pitz, W.J.; Mueller, C.J. Recent Progress in the Development of Diesel Surrogate Fuels. Prog. Energy Combust. Sci. 2011, 37, 330–350. [Google Scholar] [CrossRef]
  20. Luo, J.; Yao, M.; Liu, H.; Yang, B. Experimental and Numerical Study on Suitable Diesel Fuel Surrogates in Low Temperature Combustion Conditions. Fuel 2012, 97, 621–629. [Google Scholar] [CrossRef]
  21. Weber, J.; Won, H.W.; Peters, N. Experimental Validation of a Surrogate Fuel for Diesel; SAE International: Warrendale, PA, USA, 2007. [Google Scholar]
  22. Yu, W.; Zhao, F.; Yang, W.; Tay, K.; Xu, H. Development of an Optimization Methodology for Formulating Both Jet Fuel and Diesel Fuel Surrogates and Their Associated Skeletal Oxidation Mechanisms. Fuel 2018, 231, 361–372. [Google Scholar] [CrossRef]
  23. Chang, Y.; Jia, M.; Li, Y.; Liu, Y.; Xie, M.; Wang, H.; Reitz, R.D. Development of a Skeletal Mechanism for Diesel Surrogate Fuel by Using a Decoupling Methodology. Combust. Flame 2015, 162, 3785–3802. [Google Scholar] [CrossRef]
  24. Dong, X.; Chang, Y.; Niu, B.; Jia, M. Development of a Practical Reaction Model of Polycyclic Aromatic Hydrocarbon (PAH) Formation and Oxidation for Diesel Surrogate Fuel. Fuel 2020, 267, 117159. [Google Scholar] [CrossRef]
  25. Dagaut, P.; Cathonnet, M. The Ignition, Oxidation, and Combustion of Kerosene: A Review of Experimental and Kinetic Modeling. Prog. Energy Combust. Sci. 2006, 32, 48–92. [Google Scholar] [CrossRef]
  26. Ranzi, E.; Cavallotti, C.; Cuoci, A.; Frassoldati, A.; Pelucchi, M.; Faravelli, T. New Reaction Classes in the Kinetic Modeling of Low Temperature Oxidation of n-Alkanes. Combust. Flame 2015, 162, 1679–1691. [Google Scholar] [CrossRef]
  27. Fan, W.; Jia, M.; Chang, Y.; Xie, M. Understanding the Relationship between Cetane Number and the Ignition Delay in Shock Tubes for Different Fuels Based on a Skeletal Primary Reference Fuel (n-Hexadecane/Iso-Cetane) Mechanism. Energy Fuels 2015, 29, 3413–3427. [Google Scholar] [CrossRef]
  28. Dagaut, P.; Ristori, A.; Frassoldati, A.; Faravelli, T.; Dayma, G.; Ranzi, E. Experimental and Semi-Detailed Kinetic Modeling Study of Decalin Oxidation and Pyrolysis over a Wide Range of Conditions. Proc. Combust. Inst. 2013, 34, 289–296. [Google Scholar] [CrossRef]
  29. Man, X.; Tang, C.; Zhang, J.; Zhang, Y.; Pan, L.; Huang, Z.; Law, C.K. An Experimental and Kinetic Modeling Study of n-Propanol and i-Propanol Ignition at High Temperatures. Combust. Flame 2014, 161, 644–656. [Google Scholar] [CrossRef]
  30. Mittal, G.; Burke, S.M.; Davies, V.A.; Parajuli, B.; Metcalfe, W.K.; Curran, H.J. Autoignition of Ethanol in a Rapid Compression Machine. Combust. Flame 2014, 161, 1164–1171. [Google Scholar] [CrossRef]
  31. Chang, Y.; Jia, M.; Xiao, J.; Li, Y.; Fan, W.; Xie, M. Construction of a Skeletal Mechanism for Butanol Isomers Based on the Decoupling Methodology. Energy Convers. Manag. 2016, 128, 250–260. [Google Scholar] [CrossRef]
  32. Zhang, P.; Klippenstein, S.J.; Law, C.K. Ab Initio Kinetics for the Decomposition of Hydroxybutyl and Butoxy Radicals of n -Butanol. J. Phys. Chem. A 2013, 117, 1890–1906. [Google Scholar] [CrossRef]
  33. Sarathy, S.M.; Vranckx, S.; Yasunaga, K.; Mehl, M.; Oßwald, P.; Metcalfe, W.K.; Westbrook, C.K.; Pitz, W.J.; Kohse-Höinghaus, K.; Fernandes, R.X.; et al. A Comprehensive Chemical Kinetic Combustion Model for the Four Butanol Isomers. Combust. Flame 2012, 159, 2028–2055. [Google Scholar] [CrossRef]
  34. Tao, F.; Reitz, R.D.; Foster, D.E.; Liu, Y. Nine-Step Phenomenological Diesel Soot Model Validated over a Wide Range of Engine Conditions. Int. J. Therm. Sci. 2009, 48, 1223–1234. [Google Scholar] [CrossRef]
  35. Vishwanathan, G.; Reitz, R.D. Development of a Practical Soot Modeling Approach and Its Application to Low-Temperature Diesel Combustion. Combust. Sci. Technol. 2010, 182, 1050–1082. [Google Scholar] [CrossRef]
  36. Bergman, M.; Golovitchev, V.I. Application of Transient Temperature vs. Equivalence Ratio Emission Maps to Engine Simulations; SAE International: Warrendale, PA, USA, 2007. [Google Scholar]
  37. Vasu, S.S.; Davidson, D.F.; Hong, Z.; Vasudevan, V.; Hanson, R.K. N-Dodecane Oxidation at High-Pressures: Measurements of Ignition Delay Times and OH Concentration Time-Histories. Proc. Combust. Inst. 2009, 32, 173–180. [Google Scholar] [CrossRef]
  38. Campbell, M.F.; Davidson, D.F.; Hanson, R.K. Ignition Delay Times of Very-Low-Vapor-Pressure Biodiesel Surrogates behind Reflected Shock Waves. Fuel 2014, 126, 271–281. [Google Scholar] [CrossRef]
  39. Fieweger, K.; Blumenthal, R.; Adomeit, G. Self-Ignition of S.I. Engine Model Fuels: A Shock Tube Investigation at High Pressure. Combust. Flame 1997, 109, 599–619. [Google Scholar] [CrossRef]
  40. Davidson, D.F.; Gauthier, B.M.; Hanson, R.K. Shock Tube Ignition Measurements of Iso-Octane/Air and Toluene/Air at High Pressures. Proc. Combust. Inst. 2005, 30, 1175–1182. [Google Scholar] [CrossRef]
  41. Oehlschlaeger, M.A.; Steinberg, J.; Westbrook, C.K.; Pitz, W.J. The Autoignition of Iso-Cetane at High to Moderate Temperatures and Elevated Pressures: Shock Tube Experiments and Kinetic Modeling. Combust. Flame 2009, 156, 2165–2172. [Google Scholar] [CrossRef]
  42. Shen, H.-P.S.; Vanderover, J.; Oehlschlaeger, M.A. A Shock Tube Study of the Auto-Ignition of Toluene/Air Mixtures at High Pressures. Proc. Combust. Inst. 2009, 32, 165–172. [Google Scholar] [CrossRef]
  43. Westbrook, C.K.; Pitz, W.J.; Herbinet, O.; Curran, H.J.; Silke, E.J. A Comprehensive Detailed Chemical Kinetic Reaction Mechanism for Combustion of n-Alkane Hydrocarbons from n-Octane to n-Hexadecane. Combust. Flame 2009, 156, 181–199. [Google Scholar] [CrossRef]
  44. Gowdagiri, S.; Wang, W.; Oehlschlaeger, M.A. A Shock Tube Ignition Delay Study of Conventional Diesel Fuel and Hydroprocessed Renewable Diesel Fuel from Algal Oil. Fuel 2014, 128, 21–29. [Google Scholar] [CrossRef]
  45. Dunphy, M.P.; Simmie, J.M. High-Temperature Oxidation of Ethanol. Part 1—Ignition Delays in Shock Waves. J. Chem. Soc. Faraday Trans. 1991, 87, 1691–1696. [Google Scholar] [CrossRef]
  46. Heufer, K.A.; Fernandes, R.X.; Olivier, H.; Beeckmann, J.; Röhl, O.; Peters, N. Shock Tube Investigations of Ignition Delays of n-Butanol at Elevated Pressures between 770 and 1250K. Proc. Combust. Inst. 2011, 33, 359–366. [Google Scholar] [CrossRef]
  47. Vranckx, S.; Heufer, K.A.; Lee, C.; Olivier, H.; Schill, L.; Kopp, W.A.; Leonhard, K.; Taatjes, C.A.; Fernandes, R.X. Role of Peroxy Chemistry in the High-Pressure Ignition of n-Butanol—Experiments and Detailed Kinetic Modelling. Combust. Flame 2011, 158, 1444–1455. [Google Scholar] [CrossRef]
  48. Hui, X.; Sung, C.-J. Laminar Flame Speeds of Transportation-Relevant Hydrocarbons and Jet Fuels at Elevated Temperatures and Pressures. Fuel 2013, 109, 191–200. [Google Scholar] [CrossRef]
  49. Ji, C.; Dames, E.; Wang, Y.L.; Wang, H.; Egolfopoulos, F.N. Propagation and Extinction of Premixed C5–C12 n-Alkane Flames. Combust. Flame 2010, 157, 277–287. [Google Scholar] [CrossRef]
  50. Davis, S.G.; Law, C.K. Laminar Flame Speeds and Oxidation Kinetics of Iso-Octane-Air and n-Heptane-Air Flames. Symp. Int. Combust. 1998, 27, 521–527. [Google Scholar] [CrossRef]
  51. Johnston, R.J.; Farrell, J.T. Laminar Burning Velocities and Markstein Lengths of Aromatics at Elevated Temperature and Pressure. Proc. Combust. Inst. 2005, 30, 217–224. [Google Scholar] [CrossRef]
  52. Li, G.; Yang, W.; Tay, K.L.; Yu, W.; Chen, L. A Reduced and Robust Reaction Mechanism for Toluene and Decalin Oxidation with Polycyclic Aromatic Hydrocarbon Predictions. Fuel 2020, 259, 116233. [Google Scholar] [CrossRef]
  53. Nakamura, H.; Suzuki, S.; Tezuka, T.; Hasegawa, S.; Maruta, K. Sooting Limits and PAH Formation of n-Hexadecane and 2,2,4,4,6,8,8-Heptamethylnonane in a Micro Flow Reactor with a Controlled Temperature Profile. Proc. Combust. Inst. 2015, 35, 3397–3404. [Google Scholar] [CrossRef]
  54. Dooley, S.; Won, S.H.; Heyne, J.; Farouk, T.I.; Ju, Y.; Dryer, F.L.; Kumar, K.; Hui, X.; Sung, C.-J.; Wang, H.; et al. The Experimental Evaluation of a Methodology for Surrogate Fuel Formulation to Emulate Gas Phase Combustion Kinetic Phenomena. Combust. Flame 2012, 159, 1444–1466. [Google Scholar] [CrossRef]
  55. Li, W.; Zhang, Y.; Mei, B.; Li, Y.; Cao, C.; Zou, J.; Yang, J.; Cheng, Z. Experimental and Kinetic Modeling Study of n-Propanol and i-Propanol Combustion: Flow Reactor Pyrolysis and Laminar Flame Propagation. Combust. Flame 2019, 207, 171–185. [Google Scholar] [CrossRef]
  56. Veloo, P.S.; Wang, Y.L.; Egolfopoulos, F.N.; Westbrook, C.K. A Comparative Experimental and Computational Study of Methanol, Ethanol, and n-Butanol Flames. Combust. Flame 2010, 157, 1989–2004. [Google Scholar] [CrossRef]
  57. van Lipzig, J.P.J.; Nilsson, E.J.K.; de Goey, L.P.H.; Konnov, A.A. Laminar Burning Velocities of n-Heptane, Iso-Octane, Ethanol and Their Binary and Tertiary Mixtures. Fuel 2011, 90, 2773–2781. [Google Scholar] [CrossRef]
  58. Egolfopoulos, F.N.; Du, D.X.; Law, C.K. A Study on Ethanol Oxidation Kinetics in Laminar Premixed Flames, Flow Reactors, and Shock Tubes. Symp. Int. Combust. 1992, 24, 833–841. [Google Scholar] [CrossRef]
  59. Vancoillie, J.; Christensen, M.; Nilsson, E.J.K.; Verhelst, S.; Konnov, A.A. Temperature Dependence of the Laminar Burning Velocity of Methanol Flames. Energy Fuels 2012, 26, 1557–1564. [Google Scholar] [CrossRef]
  60. Broustail, G.; Seers, P.; Halter, F.; Moréac, G.; Mounaim-Rousselle, C. Experimental Determination of Laminar Burning Velocity for Butanol and Ethanol Iso-Octane Blends. Fuel 2011, 90, 1–6. [Google Scholar] [CrossRef]
  61. Zeng, M.; Wullenkord, J.; Graf, I.; Kohse-Höinghaus, K. Influence of Dimethyl Ether and Diethyl Ether Addition on the Flame Structure and Pollutant Formation in Premixed Iso-Octane Flames. Combust. Flame 2017, 184, 41–54. [Google Scholar] [CrossRef]
  62. Mzé-Ahmed, A.; Hadj-Ali, K.; Dagaut, P.; Dayma, G. Experimental and Modeling Study of the Oxidation Kinetics of n-Undecane and n-Dodecane in a Jet-Stirred Reactor. Energy Fuels 2012, 26, 4253–4268. [Google Scholar] [CrossRef]
  63. Dagaut, P.; Hadj-Ali, K. Chemical Kinetic Study of the Oxidation of Isocetane (2,2,4,4,6,8,8-Heptamethylnonane) in a Jet-Stirred Reactor: Experimental and Modeling. Energy Fuels 2009, 23, 2389–2395. [Google Scholar] [CrossRef]
  64. Li, Y.; Cai, J.; Zhang, L.; Yuan, T.; Zhang, K.; Qi, F. Investigation on Chemical Structures of Premixed Toluene Flames at Low Pressure. Proc. Combust. Inst. 2011, 33, 593–600. [Google Scholar] [CrossRef]
  65. Zeng, M.; Li, Y.; Yuan, W.; Li, T.; Wang, Y.; Zhou, Z.; Zhang, L.; Qi, F. Experimental and Kinetic Modeling Study of Laminar Premixed Decalin Flames. Proc. Combust. Inst. 2017, 36, 1193–1202. [Google Scholar] [CrossRef]
  66. Skeen, S.A.; Yang, B.; Jasper, A.W.; Pitz, W.J.; Hansen, N. Chemical Structures of Low-Pressure Premixed Methylcyclohexane Flames as Benchmarks for the Development of a Predictive Combustion Chemistry Model. Energy Fuels 2011, 25, 5611–5625. [Google Scholar] [CrossRef]
  67. Mati, K.; Ristori, A.; Gaïl, S.; Pengloan, G.; Dagaut, P. The Oxidation of a Diesel Fuel at 1–10 Atm: Experimental Study in a JSR and Detailed Chemical Kinetic Modeling. Proc. Combust. Inst. 2007, 31, 2939–2946. [Google Scholar] [CrossRef]
  68. Togbé, C.; Dagaut, P.; Halter, F.; Foucher, F. 2-Propanol Oxidation in a Pressurized Jet-Stirred Reactor (JSR) and Combustion Bomb: Experimental and Detailed Kinetic Modeling Study. Energy Fuels 2011, 25, 676–683. [Google Scholar] [CrossRef]
  69. Hansen, N.; Harper, M.R.; Green, W.H. High-Temperature Oxidation Chemistry of n-Butanol–Experiments in Low-Pressure Premixed Flames and Detailed Kinetic Modeling. Phys. Chem. Chem. Phys. 2011, 13, 20262. [Google Scholar] [CrossRef]
Figure 1. Diagram illustrating development of IBE–diesel combustion mechanism.
Figure 1. Diagram illustrating development of IBE–diesel combustion mechanism.
Processes 12 00995 g001
Figure 2. Oxidation routes of n-dodecane, iso-octane, and iso-cetane.
Figure 2. Oxidation routes of n-dodecane, iso-octane, and iso-cetane.
Processes 12 00995 g002
Figure 3. Oxidation routes of decalin.
Figure 3. Oxidation routes of decalin.
Processes 12 00995 g003
Figure 4. Major oxidation routes of toluene.
Figure 4. Major oxidation routes of toluene.
Processes 12 00995 g004
Figure 5. Major oxidation routes of iso-propanol, ethanol, and n-butanol.
Figure 5. Major oxidation routes of iso-propanol, ethanol, and n-butanol.
Processes 12 00995 g005
Figure 6. Measured and simulated ignition delays of n-dodecane, iso-octane, and iso-cetane.
Figure 6. Measured and simulated ignition delays of n-dodecane, iso-octane, and iso-cetane.
Processes 12 00995 g006
Figure 7. Measured and simulated ignition delays of toluene.
Figure 7. Measured and simulated ignition delays of toluene.
Processes 12 00995 g007
Figure 8. Measured and simulated ignition delays of decalin.
Figure 8. Measured and simulated ignition delays of decalin.
Processes 12 00995 g008
Figure 9. Measured and simulated ignition delays of diesel.
Figure 9. Measured and simulated ignition delays of diesel.
Processes 12 00995 g009
Figure 10. Measured and simulated ignition delays of iso-propanol.
Figure 10. Measured and simulated ignition delays of iso-propanol.
Processes 12 00995 g010
Figure 11. Measured and simulated ignition delays of ethanol and n-butanol.
Figure 11. Measured and simulated ignition delays of ethanol and n-butanol.
Processes 12 00995 g011
Figure 12. Measured and simulated laminar flame speed of diesel surrogates [27,48,49,50,51,52,53].
Figure 12. Measured and simulated laminar flame speed of diesel surrogates [27,48,49,50,51,52,53].
Processes 12 00995 g012
Figure 13. Measured and simulated laminar flame speed of IBE components.
Figure 13. Measured and simulated laminar flame speed of IBE components.
Processes 12 00995 g013
Figure 14. The measured and simulated species profile in the premixed flame of iso-octane.
Figure 14. The measured and simulated species profile in the premixed flame of iso-octane.
Processes 12 00995 g014
Figure 15. The measured and simulated species profile in the premixed flame of n-dodecane.
Figure 15. The measured and simulated species profile in the premixed flame of n-dodecane.
Processes 12 00995 g015
Figure 16. The measured and simulated species profile in the premixed flame of iso-cetane.
Figure 16. The measured and simulated species profile in the premixed flame of iso-cetane.
Processes 12 00995 g016
Figure 17. The measured and simulated species profile in the premixed flame of toluene and decalin.
Figure 17. The measured and simulated species profile in the premixed flame of toluene and decalin.
Processes 12 00995 g017
Figure 18. The measured and simulated species profile in the premixed flame of diesel.
Figure 18. The measured and simulated species profile in the premixed flame of diesel.
Processes 12 00995 g018
Figure 19. The measured and simulated species profile in the premixed flame of IBE components.
Figure 19. The measured and simulated species profile in the premixed flame of IBE components.
Processes 12 00995 g019
Figure 20. Computational grid of engine simulation at top dead center.
Figure 20. Computational grid of engine simulation at top dead center.
Processes 12 00995 g020
Figure 21. Measured and simulated in-cylinder pressure and heat release rate.
Figure 21. Measured and simulated in-cylinder pressure and heat release rate.
Processes 12 00995 g021
Table 1. Engine specifications and operation conditions.
Table 1. Engine specifications and operation conditions.
Engine SpecificationsOperation Conditions
Type4-stroke diesel engineEngine speed (rpm)1000
Bore stroke (mm)81 × 88Amount of injected fuel (mg/cycle)17.9
Displaced volume (L)2.7 Pilot start of injection (°CA)344.5
Compression ratio17.3:1Main start of injection (°CA)362.5
Diesel injection systemCommon railPilot/main injection ratio1/12
Number of injection holes6Intake valve close (°CA)210
Diameter of injection holes (mm)0.127Exhaust valve open (°CA)490
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

Ma, Y.; Zhao, S.; Zhao, J.; Fu, J.; Yuan, W. Constructing a Skeletal Iso-Propanol–Butanol–Ethanol (IBE)–Diesel Mechanism Using the Decoupling Method. Processes 2024, 12, 995. https://doi.org/10.3390/pr12050995

AMA Style

Ma Y, Zhao S, Zhao J, Fu J, Yuan W. Constructing a Skeletal Iso-Propanol–Butanol–Ethanol (IBE)–Diesel Mechanism Using the Decoupling Method. Processes. 2024; 12(5):995. https://doi.org/10.3390/pr12050995

Chicago/Turabian Style

Ma, Yi, Shaomin Zhao, Junhong Zhao, Jun Fu, and Wenhua Yuan. 2024. "Constructing a Skeletal Iso-Propanol–Butanol–Ethanol (IBE)–Diesel Mechanism Using the Decoupling Method" Processes 12, no. 5: 995. https://doi.org/10.3390/pr12050995

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