Next Article in Journal
Enhanced Saccharification of Purple Alfalfa via Sequential Pretreatment with Acidified Ethylene Glycol and Urea/NaOH
Next Article in Special Issue
Optimization Analysis of Locomotive Diesel Engine Intake System Based on Matlab-Simulink and GT-Power
Previous Article in Journal
Removal of Acetone Vapor from Air Using a Biotrickling Filter Packed with Polymeric Bioballs
Previous Article in Special Issue
Effects of Different Biodiesel-Diesel Blend Fuel on Combustion and Emission Characteristics of a Diesel Engine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Pre-Injection on Combustion and Emission Characteristics of a Diesel Engine Fueled with Diesel/Methanol/n-Butanol Blended Fuel

1
School of Mechanical and Electrical Engineering, Central South University of Forestry and Technology, Changsha 410004, China
2
School of Informational and Electrical Engineering, Hunan University of Science and Technology, Xiangtan 411201, China
*
Author to whom correspondence should be addressed.
Processes 2022, 10(1), 60; https://doi.org/10.3390/pr10010060
Submission received: 27 November 2021 / Revised: 23 December 2021 / Accepted: 24 December 2021 / Published: 28 December 2021
(This article belongs to the Special Issue Clean Combustion and Emission in Vehicle Power System)

Abstract

:
In this study, the combustion and emission characteristics of a diesel/methanol/n-butanol blended fuel engine with different pre-injection timings and pre-injection mass ratios were investigated by a computational fluid dynamics (CFD) model. The CFD model was verified by the measured results and coupled with a simplified chemical kinetics mechanism. Firstly, the corresponding three-dimensional CFD model was established by CONVERGE software and the CHEKMIN program, and a chemical kinetic mechanism containing 359 reactions and 77 species was developed. Secondly, the combustion and emission characteristics of the diesel engine with different diesel/methanol/n-butanol blended fuels were analyzed and discussed. The results showed that increases in the pre-injection timing and the pre-injection mass ratio could increase cylinder pressure and cylinder temperature and decrease soot, HC, and CO emissions. At 100% load, the maximum cylinder pressures at the start of pre-injection timing from −15 °CA to −45 °CA, were 7.71, 9.46, 9.85, 9.912, and 9.95 MPa, respectively. The maximum cylinder pressures at pre-injection fuel mass ratios from 0.1 to 0.9 were 7.98, 9.10, 9.96, 10.52, and 11.16 MPa, respectively. At 50% load, with increases of the pre-injection timing and pre-injection fuel mass ratio, the soot emission decreased by 7.30%, 9.45%, 27.70%, 66.80%, 81.80% and 11.30%, 20.03%, 71.32%, 83.80%, 93.76%, respectively, and CO emissions were reduced by 5.77%, 12.31%, 22.73%, 53.59%, 63.22% and 8.29%, 43.97%, 53.59%, 58.86%, 61.18%, respectively. However, with increases of the pre-injection timing and pre-injection mass ratio, NOx emission increased. In addition, it was found that the optimal pre-injection timing and optimal pre-injection mass ratio should be −30 °CA and 0.5, respectively. Therefore, through this study we can better understand the potential interaction of relevant parameters and propose pre-injection solutions to improve combustion and emission characteristics.

1. Introduction

Due to its high thermal efficiency, good reliability, and sufficient and stable power, the diesel engine has been favored for a long time by the fields of national defense and the military, and in power plants, transportation, engineering machinery, and agricultural machinery [1]. It has made great contributions to improvements in social development and at the economic level [2]. However, the development of the diesel engine has faced great challenges with the collection and utilization of oil resources [3]. This is mainly reflected in energy shortages caused by the massive utilization of fossil fuels and the environmental pollution caused by combustion emissions [4]. Therefore, facing the global energy crisis and environmental degradation [5], how to find renewable energy, and how to effectively reduce diesel engine emissions are the problems that need to be solved in today’s society [6]. An important method is to use alternative fuels in diesel engines [7]. The use of alternative fuels in diesel engines is considered to be an effective measure to reduce their dependence on diesel and reduce pollutant emissions [8]. Therefore, the engine parameters of various biofuels (such as biodiesel and alcohol) need to be optimized to enhance their characteristics [9]. As a result, research on exploring new renewable alternative fuels has attracted more and more attention [10,11].
Methanol and ethanol belong to low-carbon chain alcohols. As pure diesel additives, their oxygen-containing structural characteristics can improve the engine’s combustion process and reduce soot emission [12]. Compared with traditional fossil fuels, methanol has obvious advantages [13]. For example, with a high octane number and good seismic performance, the engine compression ratio can be appropriately increased to improve engine performance and reduce fuel consumption. Moreover, methanol has a large latent heat of vaporization and lower combustion temperature in the cylinder, which can reduce heat loss and NOx emission [14]. Park et al. [15] studied the effect of methanol addition on the performance and emission characteristics of diesel engines. The results showed that the cylinder pressure, cylinder temperature, and NOx emission decreased with increased methanol content. However, due to the characteristics of a low cetane number and high latent heat of evaporation of methanol, spontaneous combustion occurs easily [16]. In addition, due to the low calorific value, poor stability and immiscibility with diesel [17], the application of methanol in diesel engines is greatly hindered [18,19].
Compared with other fuels, n-butanol has a low latent heat of vaporization and ignition temperature. Thus, the cold start characteristic of the diesel engine can be greatly improved [20]. In addition, n-butanol is more resistant to water pollution and less corrosive. The most important point is that n-butanol is more miscible with diesel [21]. Moreover, the viscosity and density of n-butanol are similar to diesel [22]. Lots of experts have carried out research on n-butanol. For example, Satsangi et al. [23] studied the combustion and emission characteristic of a diesel engine fueled with four different diesel/n-butanol blends. The results showed that compared with pure diesel, the maximum cylinder pressure of diesel engine fuel with diesel/n-butanol blends increased by 5.3% and the emission characteristics of blends of diesel engine fuel with diesel/n-butanol also improved. Rakopolulos et al. [24] studied the effect of diesel/n-butanol blends on the performance and emission characteristics of a diesel engine. The results showed that the brake specific fuel consumption (BSFC) of a diesel engine fueled with a diesel/n-butanol blend increased due to lower calorific values than pure diesel. However, the brake thermal efficiency (BTE) improved due to the improved combustion caused by the n-butanol content in the diesel/n-butanol blend.
To improve engine performance, combustion and emission, this problem can be overcome by changing factors, such as the pilot fuel volume, injection timing, gaseous fuel composition and intake conditions [25]. Pre-injection is a form of multi-injection and means that a small amount of fuel is injected first, before the main injection. The main parameters, such as the pre-injection timing, pre-injection fuel mass ratio, and injection duration, will significantly affect combustion and emission characteristics [26]. In diesel engines, many researchers had studied the effects of pre-injection on engine combustion and emission characteristics. For example, Lu et al. [27] and Wang et al. [28] investigated the effects of the main injection mode, injection ratio, and injection timing on the combustion and emission characteristics of a high-load, low-speed diesel engine. The results showed that the advanced pre-injection timing and increased pre-injection fuel mass ratio could reduce knock and NOx emission. Similarly, Liu et al. [29] studied the effects of injection timing on the performance and exhaust emissions of the engine diesel/methanol compound combustion (DMCC) mode. The results showed that under the DMCC model, the peak cylinder pressure and maximum heat release rate in the cylinder were increased, and the intake air temperature, NOx, and soot emissions were reduced. Liu et al. [30] investigated the effects of each parameter on the emission characteristics of a diesel engine fueled with an n-butanol/diesel blended fuel under two combustion strategies, early injection partial premixed combustion (PPC) and pre-injection PPC. The results showed that the premixed combustion ratio and the average diameter of particulate matter decreased with an increasing pre-injection ratio. Bhowmick et al. [31] investigated the effect of different injection strategies on a common rail direct injection engine. The results showed that the 10% pilot fuel and 90% main injection strategy (B10@P10-M90) outperformed the pure diesel fuel among all the injection strategies. Compared to all the injection strategies, the B10@P10-M90 had the highest efficiency of 35.8% and the lowest fuel consumption of 0.25 kg/(kW·h). The carbon monoxide (CO) and hydrocarbon (HC) emissions were lower than all the tested emissions of all the samples tested.
With the development of computer science and information technology, the combustion chemical kinetic mechanism has been further expanded and examined in depth. Moreover, in the manufacturing industry, design, experimentation, and prototyping are very difficult and expensive [32]. Thus, the simulation method has been applied in the internal combustion engine field [33]. Simulations can not only greatly save in the cost, but also greatly shorten the research cycle [34]. The numerical simulation has been widely accepted as a good way to study diesel engines. This is because the three-dimensional computational fluid dynamics (CFD) simulation method combined with the chemical kinetic mechanism can accurately predict the combustion process [35]. Many researchers had studied n-butanol/diesel blends by the CFD simulation method. Gong et al. [36] used AVL-Fire to simulate the effects of injection and ignition timing on engine combustion and the in-cylinder formaldehyde and unburned methanol emissions during the cold start of a methanol engine. Zang et al. [37] used an improved KIVA−3 V code coupled with the CHEMKIN solver to simulate the combustion and emission of a diesel/methanol dual fuel (DMDF) engine. They found that the CFD method could be considered as an effective method and could accurately predict the cylinder combustion process.
In conclusion, a large number of researchers have used CFD software to study the effects of diesel/methanol and diesel/n-butanol blended fuels on engine combustion and emission characteristics, and have further optimized engine combustion and emission combined, using a pre-injection strategy. However, very few have used CONVERGE Studio software to simulate engine combustion chambers with diesel/methanol/n-butanol blends to investigate the effects of pre-injection timing and the pre-injection fuel mass ratio on engine combustion and emission characteristics.
As mentioned above, the behaviors of diesel/methanol/n-butanol blended fuels are different due to the different thermos-physical properties. In this paper, CONVERGE software combined with the Chemkin code was employed to simulate the cylinder combustion process. Firstly, according a four-stroke diesel engine, the CFD model was developed and then validated by the experimental results at different loads. Then, the combustion processes of diesel/methanol/n-butanol blends were simulated and compared. Finally, the effects of the pre-injection timing and pre-injection mass ratio on the combustion and emission characteristics of a diesel engine fueled with diesel/methanol/n-butanol blends were studied. Therefore, this study helps to optimize the combustion and emission characteristics of diesel/methanol/n-butanol blends for diesel engines, reducing the dependence on petroleum fuels, improving diesel engine performance, and reducing diesel engine pollutant emissions.

2. Materials and Methods

In this study, the numerical simulations were performed by using CONVERGE software combined with a detailed chemical kinetic mechanism. A series of numerical sub-models were used to simulate the combustion process and pollutant formation mechanism of the diesel/methanol/n-butanol (DMB) blended fuel engine.

2.1. Turbulence Model

The RNG k-ε turbulence model was used for the calculations because this turbulence model fully considers the change of gas density in the wall boundary layer [38] and the influence of the vortex on turbulence and improves the accuracy of vortex flow [39]. Moreover, the model has low requirements for the grid quality near the wall. The principle is to introduce turbulent kinetic energy k and a turbulent dissipation rate ε based on the Navier-Stokes equation by using the basic idea of RNG. The expression is as follows [40]:
( ρ k t ) t + d i v ( ρ k U ) = d i v [ α k μ e f f grad   k t ] + τ i j S i j ρ ε
( ρ ε ) t + d i v ( ρ ε U ) = d i v [ a ε μ e f f grad   ε ] + C 1 ε * ε k t τ i j S i j C 2 ε ρ ε 2 k t
with,
τ i j = ρ u i u j ¯ = 2 μ t S i j 2 3 ρ k t δ i j
and
μ e f f = μ + μ t μ t = ρ C μ k t 2 ε
C 1 ε * = C 1 ε η ( 1 η η 0 ) 1 + β η 3 η = k t ε 2 S i j S i j
where ρ is the density, g/cm3; kt is the turbulent kinetic energy, m2/s2; U is the fluid velocity, m/s; Sij is the turbulent source term; ε is the turbulent dissipation rate; μ is the molecular viscosity, Pa·s. The Kronecker delta δij is given by δij = 1 if i = j and δij = 0 if Ij. C1ε = 1.42, C2ε = 1.68, αk = αε = 1.39, Cμ = 0.0845, η0 = 4.377, and β = 0.012.

2.2. Combustion Model

In addition, the SAGE model was chosen in this study to simulate the combustion process of a DMB blended fuel engine with a chemical kinetic mechanism including 77 species and 359 reactions.
According to the description of Turns [41], a multistep chemical reaction mechanism can be written as follows:
m = 1 M u m , i y m m = 1 M u m , i y m   for   i = 1 , 2 , I
where u m , i is the stoichiometric coefficients for the reactants for species m and reaction i; u m , i is the stoichiometric coefficients for the products for species m and reaction i; I is the total number of reactions; and ym is the chemical symbol for the species m. The net productivity of species m is
ω ˙ m = i = 1 I u m , i q i   for   m = 1 , 2 , M
where M is the total number of species and
u m , i = u m , i u m , i
The reaction rate parameter qi of the ith reaction is
q i = k i , f m = 1 M [ X m ] u m , i k i , r m = 1 M [ X m ] u m , i
where Xm is the molar concentration of species m, mol/L; ki,f and ki,r are the forward and reverse rate coefficients for reaction i. In SAGE, the forward rate coefficient is expressed in Arrhenius form as
k i , f = A i T β i exp ( E i R T )
where Ai is the pre-exponential factor; βi is the temperature exponent; Ei is the activation energy, cal/mol; R is the ideal gas constant; and T is the temperature, K. In addition, the reverse rate coefficient can either be specified in an analogous fashion as in Equation (11), or calculated from the equilibrium coefficient Ki,c as
K i , c = k i , f k i , r
The equilibrium coefficient Ki,c is determined by thermodynamic properties
K i , c = K i , p ( P a t m R T ) m = 1 M u m , i
where Patm is the atmospheric pressure, Pa. The equilibrium constant Ki,p is obtained via
K i , p = exp ( Δ S i 1 R Δ H i 1 R T )
The Δ refers to the change that occurs in passing completely from reactants to products in the ith reaction, specifically,
Δ S i 1 R = m = 1 M u m , i S m 1 R
Δ H i 1 R T = m = 1 M u m , i H m 1 R T
where S denotes entropy, J/(mol·K); and H denotes enthalpy, kJ/mol.

2.3. Spray Model

Based on the physical characteristics of the fuel spray and blending process, the Kelvin-Helmholtz and Rayleigh-Taylor (KH-RT) model is employed to simulate the spray breakup process of fuel. In this model, the KH mechanism and the RT mechanism were used to simulate the primary and secondary crushing, respectively [42].
The following formula can express the radius change rate of droplets in the KH model:
r = B o Λ K H
v n = C 1 Λ K H Ω K H
τ K H = 3.726 B b r i Λ K H Ω K H
where r is the droplet size, mm; Bo is the model size constant; ΛKH is the wavelength of the fastest growing wave on the surface of the initial liquid column, mm; vn is the model velocity constant; C1 is the model constant; ΩKH is the wave growth rate of the fastest growing wave on the surface of the initial liquid column, mm/s; τKH is the breakup time, s; Bb is the model breakup time constant; and ri is the initial droplet radius, mm.
The RT model obtains the wavelength (ΛRT) and the wave growth rate (ΩRT) by solving the discrete equation:
Ω R T = ( 2 3 3 σ s [ g t ( ρ l ρ g ) ] 3 / 2 ρ l + ρ g ) 1 / 2
where gt is the acceleration in the direction of motion, m/s2; ρl is the density of the liquid phase, g/cm3; ρg is the density of the gas phase, g/cm3; and σs is the surface tension coefficient.
The wavelength and wave number (KRT) are:
Λ R T = 2 π C R T K R T
K R T = ( g t ( ρ l ρ g ) 3 σ s )
where CRT is a parameter that can be adjusted through the measurement results related to the nozzle conditions. In this paper, it is taken as 0.1.
The droplet breaking time scale and new droplet radius (rnew) are:
τ R T = C τ Ω R T
r n e w = π C R T K R T
where Cτ is the model breakup time constant.
The RT breakup length (Lb) equation is given by [43]:
L b = C b l ρ l ρ g d o
where Cbl is the RT model breakup length constant; and do is the nozzle diameter, mm.
Due to the low energy substitution rate and less collision with the wall, the rebound/slide model [43] was more suitable for the interaction between droplets and solid surfaces. The model includes two impact regimes, rebound and slide, which are based on the Weber number Wei of the incoming drop at impact.
W e i = ρ f V n 2 d 0 σ f
where ρf is the fluid density (kg/m3); Vn is the velocity component normal to the surface, m/s; d0 is the characteristic length, m; σf is the surface tension coefficient of the fluid, N/m.
If Wei is less than 80 (rebound regime), the drop rebounds elastically with a normal velocity given by [43] as
V n , o = V n , i W e o W e i
where Vn,o is the velocity component normal to the surface of the outgoing drop, m/s; Vn,i is the velocity component normal to the surface of the incoming drop, m/s. The Weber number of the outgoing drop Weo is obtained from:
W e o = 0.678 W e i exp ( 0.04415 W e i )
If Wei is greater than 80, the jet model of Naber and Reitz [44] is used to update the drop velocity.
Moreover, the Frossling model [45] was used to predict droplet evaporation. The Frossling correlation is
d r i d t = α s p r a y ρ g D 2 ρ l r i B d S h d
where αspray is the scaling factor for the mass transfer coefficient; D is the mass diffusivity of liquid vapor in air, m2/s. It defines Bd as:
B d = Y 1 * Y 1 1 Y 1 *
where Y1* is the vapor mass fraction at the drop’s surface, %; Y1 is the vapor mass fraction, %; and Shd is the Sherwood number given by:
S h d = ( 2.0 + 0.6 Re d 1 / 2 S c 1 / 3 ) ln ( 1 + B d ) B d
where
Re d = ρ g | u i + u i v i | d μ a i r
where ui and ui are the local mean and turbulent fluctuating gas velocities, respectively, m/s; vi is the velocity of the droplet, m/s; d is the drop diameter, mm; and μair is the air viscosity which is evaluated at the temperature T given by [45]
T = T g a s + 2 T d 3
where Tgas is the gas temperature, K; and Td is the drop temperature, K.
Furthermore, the Sc is the Schmidt number of air given by
S c = μ a i r ρ g a s D
In addition the mass diffusivity of liquid vapor in air D is determined from the correlation:
ρ g D = 1.293 D 0 ( T / 273 ) n 0 1
where D0 and n0 are model constants.
Y 1 * = M W C n H 2 m M W C n H 2 m + M W m i x ( p g a s p v 1 )
where M W C n H 2 m is the molecular weight of the CnH2m, Da; MWmix is the molecular weight of the mixture (not including vapor from the liquid species), Da; pgas is the gas pressure, Pa; and pv is the vapor pressure at the current droplet temperature, Pa.

2.4. Emission Model

The formation of NOx was modeled using the Extended Zeldovich mechanism proposed by Heywood [46]. This model can better predict the formation of NO. The expression is as follows:
O + N 2 NO + N
O 2 + N NO + O
OH + N NO + H
The rate constants for the reactions in Equations (36)–(38) are
k 1 , f = 7.6 × 10 13 exp ( 38000 T )
k 1 , r = 1.6 × 10 13
k 2 , f = 6.4 × 10 9 T exp ( 3150 T )
k 2 , r = 1.5 × 10 9 T exp ( 19500 T )
k 3 , f = 4.1 × 10 13
k 3 , f = 2.0 × 10 14 exp ( 23650 T )
where k is the rate constant, cm3/(mol·s); the subscript f denotes a forward reaction, and the subscript r denotes a reverse reaction.
The Hiroyasu-NSC soot model was used to predict soot emission [47]. The expression is as follows:
d m s d t = d m s f d t d m s o d t
d m s f d t = A f m f P 0.5 exp ( E s f R T )
d m s o d t = A o m s P o 2 P P 1.8 exp ( E s o R u T )
where ms is the total amount of soot, kg; msf is the amount of soot generation, kg; mso is the amount of soot oxidation, kg; Af is the constant; mf is the fuel evaporation, kg; P is the cylinder pressure, Pa; Esf represents the activation energy in the process of formation; Ao is the constant; Po2 is the partial pressure of oxygen, Pa; Eso represents the activation energy in the process of oxidation, kJ; and Ru is the gas constant, cal/(K·g·mol).

2.5. Boundary Condition

In this paper, a four-cylinder, four-stroke diesel engine was used for experiments. The fuel injection quantity was obtained and adjusted by the electronic control unit of the diesel engine. The diesel engine specifications and boundary conditions are shown in Table 1. In addition, the friction work of the diesel engine was obtained by an empirical equation. The boundary conditions were obtained by the experiment. However, the boundary conditions which were not measured by the experiment were calculated by the validated AVL-BOOST model.

2.6. Computational Mesh

In this study, the engine’s combustion chamber was modeled and calculated by using CONVERGE software to simulate the combustion process of a DMB blended fuel engine. Since the fuel injection nozzle was located on the central axis of the cylinder block close to the cylinder head, the eight fuel injection nozzle holes were symmetrically distributed. In order to reduce the calculation time, the cylinder block model was simplified to a 45° sector of a single fuel nozzle hole. The geometry of the eighth combustor is shown in Figure 1. All the meshes have fine mesh near the piston crevice, fuel spray path and injector nozzle regions.
In addition, the calculation time and accuracy of the CFD model are determined by the grid size [48]. Figure 2 shows the cylinder pressure generated by the grids of the DMB blended fuel at 100% load. The results showed that there was no significant difference between 1.2 mm and 1.4 mm grids. Therefore, due to a higher accuracy and lower calculation time, the best 1.4 mm grid was used to simulate the combustion process [49].

2.7. Engine Bench Test

This study investigated the combustion and emission characteristics of a diesel engine at a speed of 2000 rpm. Before the test, the engine was first warmed up in pure diesel fuel mode until the coolant temperature was between 65 and 85 °C. Then, all tests were arranged at an ambient temperature of about 25 °C. In addition, the intake air temperature after the intercooler was kept below 40 °C. When the engine was running in a stable condition in each test case, the in-cylinder pressure was recorded for 150 consecutive engine cycles and then averaged to ensure more accurate test results. At the same time, the exhaust gas from the engine was sampled using a probe, and then transmitted to an exhaust gas analyzer to obtain the composition of the exhaust gas, including soot, NOx, HC, and CO.
Detailed physical properties of the fuel are shown in Table 2 [50,51,52]. The main specifications are listed in Table 1. The schematic diagram of the engine test is shown in Figure 3. The detailed specifications of the main measuring instruments are shown in Table 3.

2.8. Uncertainty Analysis

Generally, the experimental measurement results obtained have some errors and uncertainties [53]. The uncertainty of the experimental results is caused by the selection, observation, and calibration of sensors. The measurement results are used to calculate the required experimental results. The uncertainty percentage of parameters such as BSFC, BTE, and NOx can be obtained from Equation (48). The R in Equation (49) is a function of the independent variables I1, I2,…, In. Let u1, u2, …, un be the uncertainty in the independent variable and UR be the uncertainty in the result.
U R = { [ ( R / I 1 ) u 1 ] 2 + [ ( R / I 2 ) u 2 ] 2 + + [ ( R / I n ) u n ] 2 } 1 / 2
R = { I 1 , I 2 , I 3 , I n }
The measuring range and accuracy of the measuring equipment used in this paper are shown in Table 4. The following equation shows the calculated total experimental uncertainty.
Total uncertainty of the experiment =
Square root of [(uncertainty of pressure sensor)2 + (uncertainty of HC emission)2
+ (uncertainty of NOx emission)2 + (uncertainty of CO emission)2
+ (uncertainty of Soot emission)2]
= Square root of [(0.5%)2 + (0.11%)2 + (0.53%)2 + (0.32%)2 + (2.8%)2] = 2.913%

2.9. Model Validation

In order to verify the accuracy of the model, experiments were carried out. This paper validates the cylinder pressure (see Figure 4a) and heat release rate (see Figure 4b) of the combustion process and the NOx and soot emissions of the emission process for the D70M20B10 (70% diesel, 20% methanol and 10% n-butanol by vol) blended fuel at 50% and 100% loads. Figure 4 shows the comparisons of the cylinder pressure and heat release rate at 100% and 50% loads, respectively. It was found that the experimental results were in good agreement with the simulation results. The maximum error was less than 4%. Thus, the results showed that the model could accurately simulate the in-cylinder combustion process. Figure 5 shows the simulation and experimental soot and NOx emissions at 100% and 50% loads, respectively. It was found that the soot and NOx emission trends were similar to the experiment. Therefore, this model can simulate the effect of pre-injection on combustion and emission characteristics.

3. Results and Discussions

To investigate the effects of the pre-injection timing and pre-injection mass ratio on the combustion and emission characteristics of the diesel engine, the CFD model was employed to simulate the combustion process. The specific pre-injection strategy parameters are shown in Table 5. The parameters defined in the pre-injection study are as follows. The start of pre-injection is defined as SOPI; the pre-injection mass ratio is defined as Mpre (where Mpre = mpre/(mmain + mpre)); the start of the main injection is defined as SOMI; the main injection mass is defined as Mmain (where Mmain = mmain/(mmain + mpre)). The simulation results were then presented and analyzed in detail. Figure 6 shows the layout of the injector location in the CONVERGE environment.

3.1. Combustion Analysis

3.1.1. Cylinder Pressure

Figure 7 shows comparisons of the cylinder pressure under different load conditions. Figure 7a,c shows that the cylinder pressure increases with the increase of SOPI. At 100% load, the maximum cylinder pressure at SOPI = −15, −25, −30, −35 and −45 °CA were 7.71, 9.46, 9.85, 9.912 and 9.95 MPa, respectively. The main reason is that the increase of SOPI leads to an advance of the combustion starting point, resulting in an increase of cylinder pressure. In addition, with the increase of SOPI, the turbulent kinetic energy in the combustion chamber increases, which increases the evaporation speed of the fuel and improves the in-cylinder combustion. Thus, the cylinder pressure increases. As shown in Figure 7b,d, the cylinder pressure increases with the increase of Mpre. At 100% load, the maximum cylinder pressures at Mpre = 0.1, 0.3, 0.5, 0.7 and 0.9 were 7.98, 9.10, 9.96, 10.52 and 11.16 MPa, respectively. This is because when SOPI remains unchanged, with the increase of Mpre, more fuel enters the cylinder earlier to participate in combustion, and more fuel is compressed during the compression stroke, increasing the cylinder pressure. It can be seen that with the increase of SOPI and Mpre, the cylinder pressure of the diesel engine can be increased and the performance of the diesel engine can be improved.

3.1.2. Heat Release Rate (HRR)

Figure 8 shows comparisons of the HRR under different load conditions. As shown in Figure 8a,c, in all cases, the ignition delay was earliest at SOPI = −15 °CA, followed by −25, −30, −35, and −45 °CA. In addition, an increase of SOPI leads to an increase of the HRR peak and changes the HRR curve from a single peak to a double peak. This is because the increase of SOPI makes it easier to form a low concentration and homogeneous combustible mixture. This part of the mixture fails to fully release heat in the premixing stage, delaying the unburned blended fuel for the main injection stage and increasing the amount of fuel added to burn in the main injection stage. Therefore, the heat release rate of the main injection increases. As shown in Figure 8b,d, in all cases, the ignition delay was earliest at Mpre = 0.9, followed by 0.7, 0.5, 0.3, and 0.1. In addition, the heat release rate increases with the increase of Mpre. Due to the increase of Mpre, more fuel is compressed in the compression stroke, which accelerates the increase of the blended fuel temperature in the cylinder and improves the atomization and evaporation of the main injection diesel. Therefore, increasing HRR can significantly improve engine combustion characteristics.

3.1.3. Cylinder Temperature

Figure 9 shows the comparisons of in-cylinder temperature under different load conditions. It can be seen that the cylinder temperature increases with the increase of SOPI and Mpre. At 50% load, the maximum cylinder temperature at SOPI = −15, −25, −30, −35 and −45 °CA were 953.20, 1005.23, 1042.02, 1091.15 and 1102.46 K, respectively; the maximum cylinder temperatures at Mpre = 0.1, 0.3, 0.5, 0.7 and 0.9 were 984.04, 1062.51, 1091.16, 1134.54 and 1180.21 K, respectively. This is mainly because with the increase of SOPI and Mpre, the ignition delay period of the combustion process is prolonged, and the fuel is fully blended, which increases the combustion speed of the fuel, resulting in more heat release, thus increasing the temperature in the cylinder. In addition, with the increase of SOPI and Mpre, the cold flame effect is enhanced, and the temperature in the cylinder is increased.
Figure 10 shows the temperature distribution field under different load conditions. It can be seen that when the pre-injection strategy is adopted, the high-temperature area increases with the increase of SOPI and Mpre. When TDC = −5 °CA, the temperature of the newly injected fuel was low, which cooled the injection area of the combustion chamber. When TDC = 1 °CA, the temperature in the cylinder increases due to fuel combustion, especially at the front end of the fuel beam, where combustion occurs first. When TDC = 7 °CA, the temperature in the cylinder continues to increase with the progress of combustion, and the maximum temperature in the cylinder increases with the increase of SOPI. In addition, when SOPI = −15 °CA or −25 °CA and Mpre = 0.1 or 0.3, the temperature distribution in the cylinder was uneven. At SOPI = −30, −35, and −45 °CA and Mpre = 0.5, 0.7, and 0.9, the combustion conditions were better because of a longer fuel ignition delay and sufficient fuel blending. Thus, the in-cylinder temperature was also more evenly distributed. Therefore, increasing SOPI and Mpre within a certain range are more conducive to fuel combustion.
In addition, the generation of NOx in the cylinder is closely related to the temperature. If the temperature in the cylinder is too high it will increase the emission of NOx. Therefore, SOPI and Mpre should be reasonably selected to reduce the emission.

3.2. Emissions Analysis

3.2.1. Soot Emission

Figure 11 shows the comparison of soot changes under different load conditions. As shown in Figure 11a,c, with the increase of SOPI, the soot emission shows a downward trend. At 50% load, the soot emission at SOPI = −15, −25, −30, −35 and −45 °CA were decreased by 7.30%, 9.45%, 27.70%, 66.80% and 81.80%, respectively. Moreover, with the increase of Mpre, soot emission decreases. At 50% load, the soot emission at Mpre = 0.1, 0.3, 0.5, 0.7 and 0.9 were decreased by 11.30%, 20.03%, 71.32%, 83.80% and 93.76%, respectively. The main reason is that with the increase of SOPI and Mpre, the temperature in the combustion chamber increases, and the cold flame effect caused by pre-injection combustion is enhanced, resulting in the main injection combustion conditions being improved and so the combustion speed is accelerated. The formation condition of soot is a high temperature and anoxic environment, which is the product of an incomplete combustion of fuel and has a trade-off relationship with NOx [54]. With the increase of SOPI and Mpre, the area of an equivalent ratio in the combustion chamber decreases, the in-cylinder combustion conditions are improved to a certain extent, the combustion is more complete, and the formation of soot decreases. However, in Figure 10c,d, the soot emission firstly increases and then decreases. The reason is that at SOPI = −15 °CA and Mpre = 0.1, the cylinder pressure and cylinder temperature are low (as shown in Figure 7 and Figure 9), resulting in incomplete combustion of the fuel, which increases soot emission. In summary, a reasonable SOPI and Mpre can decrease soot emission.

3.2.2. NOx Emission

Figure 12 shows the comparison of NOx changes under different load conditions. As shown in the figure, NOx emission increases with the increase of SOPI and Mpre. In all cases, the NOx emission was highest at SOPI = −45 °CA and Mpre = 0.9, followed by SOPI = −25, −30, −35, −45 °CA and Mpre = 0.7, 0.5, 0.3, 0.1. The main reason is that when SOPI increases, the temperature in the combustion chamber increases, and the cold flame effect caused by pre-injection combustion increases, resulting in the main injection combustion conditions being improved, a faster combustion speed, and a higher combustion temperature. When Mpre increases, the blended fuel in the combustion chamber increases, and the fuel blends more fully, increasing the cylinder temperature. Moreover, NOx is produced under the conditions of a high-temperature oxygen-enriched environment and high-temperature duration. Therefore, when the speed remains unchanged, the temperature in the cylinder increases, which will increase NOx emission. It can be seen that an increase of SOPI and Mpre is not good for NOx emission, and a reasonable SOPI and Mpre can improve the performance of the diesel engine. However, diesel engine emissions should also be considered at the same time.
Figure 13 shows the NOx distribution under different conditions. As shown in the figure, at TDC = −5 °CA, only a small amount of NOx is formed due to in-cylinder combustion and the effect of waste heat in the cylinder. When TDC = 1 °CA, NOx is mainly generated in the high-temperature area at the front of the fuel beam because the formation of NOx is affected by high temperatures. When TDC = 7 °CA, with the progress of combustion, the high-temperature area gradually diffuses into the whole cylinder, and the NOx formation area increases and diffuses into the whole combustion chamber. As shown in Figure 12 and Figure 13, the NOx distribution area of the combustion chamber is consistent with the temperature, indicating that the temperature is an important factor affecting NOx.

3.2.3. HC Emission

Figure 14 shows the change of HC under different load conditions. It can be seen that with SOPI increases, HC emissions firstly increase and then decrease. At 50% load, the HC emission at SOPI = −15, −25, −30, −35 and −45 °CA changed by +42.81%, −7.94%, −30.22%, −35.75% and −41.24%, respectively. As shown in Figure 7 and Figure 9, when SOPI = −15 °CA, the cylinder pressure and cylinder temperature are low, resulting in incomplete combustion of the fuel, which increases HC emissions. With an increase of SOPI, the ignition delay of the pilot diesel increases, and more combustible blended fuel can be obtained. The combustion has more ignition points and ignition energy to improve the combustion, resulting in an increase of temperature and pressure in the cylinder, and the fuel combustion is more complete. At the same time, the earlier injection moment will make the combustion start earlier, and the HC formed by combustion will have adequate time for oxidation reactions so that HC emissions will be decreased. In addition, with the increase of Mpre, HC emission increases first and then decreases. At 50% load, the HC emission at Mpre = 0.1, 0.3, 0.5, 0.7 and 0.9 changed by +12.71%, −28.02%, −36.74%, −59.43% and −82.71%, respectively. When Mpre = 0.1, less fuel enters the cylinder. Moreover, as shown in Figure 7 and Figure 9, the cylinder pressure and cylinder temperature are low, resulting in incomplete combustion of the fuel, which increases HC emissions. With Mpre continuing to increase, the HC emissions decrease. This may be because as the fuel increases, it makes the combustion more complete. The earlier SOPI and larger Mpre can significantly decrease HC emissions, which is important for the diesel engine.

3.2.4. CO Emission

CO emission is mainly caused by incomplete combustion. In addition, carbon dioxide (CO2) cracking under a high-temperature environment in the cylinder is another important reason [55]. Figure 15 shows the change of CO percentage under different load conditions. It can be seen that CO emissions decrease with the increase of SOPI and Mpre. Taking 50% load as an example, with the increase of SOPI, CO emission changed by −5.77%, −12.31%, −22.73%, −53.59%, and −63.22%, respectively. With the increase of Mpre, CO emission changed by −8.29%, −43.97%, −53.59%, −58.86%, and −61.18%, respectively. This may be due to the increase of SOPI, so the time for evaporation and fuel blending are enough to completely burn it. In addition, due to the increase of Mpre, more fuel is provided for combustion. Therefore, CO emission is decreased. In addition, a larger SOPI and Mpre will prolong the blending time of fuel and air to decrease CO emission. It can be seen that increases of SOPI and Mpre can significantly decrease the emission of CO.

4. Conclusions

Due to the global energy crisis [56,57,58,59,60] and environmental problems [61,62,63,64,65,66], the development of diesel engines is also facing great challenges. In this paper, it is studied the effects of pre-injection timing and pre-injection mass ratio on the combustion and emission characteristics of a diesel/methanol/n-butanol blended fuel engine using computational fluid dynamics and a chemical dynamics model. The 3-D CFD simulation model was established and calibrated based on the test data under the two typical working conditions [67,68]. The intermediate process of combustion and emission can be easily obtained and used to explain the changes of combustion and emission parameters through the simplified chemical kinetic model. Through the analysis of experimental data and simulation results, the following conclusions are drawn:
(1)
When the pre-injection timing changes from −15 °CA to −45 °CA, the engine cylinder pressure increases, and the cylinder temperature increases. The increase in SOPI leads to an earlier start of combustion, which increases the evaporation rate of the fuel and improves in-cylinder combustion.
(2)
When the pre-injection fuel mass ratio changes from 0.1 to 0.9, the engine cylinder pressure increases and the cylinder temperature increases. This is because with the increase of Mpre, more fuel enters the cylinder in advance to participate in combustion, and more fuel is compressed in the compression stroke, resulting in more heat release, thus increasing cylinder pressure and cylinder temperature.
(3)
With the increase of SOPI, soot, CO and HC emissions are decreased. This is because with the increase of SOPI, more combustible blended fuels can be obtained, and the time for fuel evaporation and fuel blending is enough to make it burn completely.
(4)
With the increase of Mpre, soot, CO and HC emissions are reduced. With the increase of Mpre, more fuel is used to participate in combustion, making the combustion more complete.
(5)
NOx emission increases with SOPI and Mpre. NOx is produced under high-temperature oxygen-rich environments and high-temperature duration conditions. Therefore, the temperature in the cylinder increases, which will increase NOx emission.
From the present work, it can be concluded that the diesel engine fueled with diesel/methanol/n-butanol blends, combined with a pre-injection strategy can significantly improve diesel combustion performance and decrease emissions. Therefore, to balance the emissions of NOx, HC, CO, and soot, the optimal pre-injection timing and the optimal pre-injection mass ratio should be SOPI = −30 °CA and Mpre = 0.5.

Author Contributions

Conceptualization, Z.W. and L.L.; software, Z.W.; formal analysis, Z.W. and L.L.; investigation, Z.W. and L.L.; resources, L.L.; writing—original draft preparation, Z.W.; writing—review and editing, Z.W. and L.L.; supervision, L.L.; funding acquisition, Z.W. and L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the key research and development project of Hunan Province (2021NK2023) and the Special funds for the construction of the innovative Hunan Province (2019NK2022).

Data Availability Statement

All data used to support the findings of this study are included within the article.

Conflicts of Interest

The authors declare that they have no conflict of interest regarding the publication of this paper.

References

  1. Viswanathan, K.; Subramanian, T.; Ashok, B.; Varuvel, E.; Azad, K. Experimental investigation of pomegranate oil methyl ester in ceramic coated engine at different operating condition in direct injection diesel engine with energy and exergy analysis. Energy Convers. Manag. 2020, 205, 112334. [Google Scholar]
  2. Tan, D.; Chen, Z.; Li, J.; Luo, J.; Yang, D.; Cui, S.; Zhang, Z. Effects of Swirl and Boiling Heat Transfer on the Performance Enhancement and Emission Reduction for a Medium Diesel Engine Fueled with Biodiesel. Processes 2021, 9, 568. [Google Scholar] [CrossRef]
  3. Zhang, Z.; Ye, J.; Tan, D.; Feng, Z.; Luo, J.; Tan, Y.; Huang, Y. The effects of Fe2O3 based DOC and SCR catalyst on the combustion and emission characteristics of a diesel engine fueled with biodiesel. Fuel 2021, 290, 120039. [Google Scholar] [CrossRef]
  4. Nanthagopal, K.; Ashok, B.; Tamilarasu, A.; Johny, A.; Mohan, A. Influence on the effect of zinc oxide and titanium dioxide nanoparticles as an additive with Calophyllum inophyllum methyl ester in a CI engine. Energy Convers. Manag. 2017, 146, 8–19. [Google Scholar] [CrossRef]
  5. E, J.; Zhang, Z.; Chen, J.; Pham, M.; Zhao, X.; Peng, Q.; Zhang, B.; Yin, Z. Performance and emission evaluation of a marine diesel engine fueled by water biodiesel-diesel emulsion blends with a fuel additive of a cerium oxide nanoparticle. Energy Convers. Manag. 2018, 169, 194–205. [Google Scholar] [CrossRef]
  6. Zhang, Z.; Chen, J.; Zhu, H.; Zhao, X.; Han, D.; Zuo, W.; Peng, Q.; Gong, J.; Yin, Z. Effects of low-level water addition on spray, combustion and emission characteristics of a medium speed diesel engine fueled with biodiesel fuel. Fuel 2019, 239, 245–262. [Google Scholar] [CrossRef]
  7. E, J.; Zhao, X.; Qiu, L.; Wei, K.; Zhang, Z.; Deng, Y.; Han, D.; Liu, G. Experimental investigation on performance and economy characteristics of a diesel engine with variable nozzle turbocharger and its application in urban bus. Energy Convers. Manag. 2019, 193, 149–161. [Google Scholar] [CrossRef]
  8. Zhang, Z.; E, J.; Chen, J.; Zhao, X.; Zhang, B.; Deng, Y.; Peng, Q.; Yin, Z. Effects of boiling heat transfer on the performance enhancement of a medium speed diesel engine fueled with diesel and rapeseed methyl ester. Appl. Therm. Eng. 2020, 169, 114984. [Google Scholar] [CrossRef]
  9. Saiteja, P.; Ashok, B. A critical insight review on homogeneous charge compression ignition engine characteristics powered by biofuels. Fuel 2021, 285, 119202. [Google Scholar] [CrossRef]
  10. Viswanathan, K.; Ashok, B.; Nanthagopal, K.; Subramanian, T.; Varuvel, E. Investigation of Novel Pistacia khinjuk Biodiesel in DI Diesel Engine with Post Combustion Capture System. Appl. Therm. Eng. 2019, 159, 113969. [Google Scholar]
  11. Viswanathan, K.; Subramanian, T.; Varuvel, E.; Ashok, B.; Nanthagopal, K.; Ong, H.C.; Rajasekar, V. Simultaneous reduction of NOx and smoke emissions with low viscous biofuel in low heat rejection engine using selective catalytic reduction technique. Fuel 2019, 255, 115854. [Google Scholar]
  12. 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]
  13. Hasan, A.O.; Osman, A.I.; Al-Muhtaseb, A.H.; Al-Rawashdeh, H.; Abu-jrai, A.; Ahmad, R.; Gomaa, M.R.; Deka, T.J.; Rooney, D.W. An experimental study of engine characteristics and tailpipe emissions from modern DI diesel engine fuelled with methanol/diesel blends. Fuel Process. Technol. 2021, 220, 106901. [Google Scholar] [CrossRef]
  14. Li, Z.; Wang, Y.; Yin, Z.; Gao, Z.; Wang, Y.; Zhen, X. Parametric study of a single-channel diesel/methanol dual-fuel injector on a diesel engine fueled with directly injected methanol and pilot diesel. Fuel 2021, 302, 121156. [Google Scholar] [CrossRef]
  15. Park, S.; Jungkeun, C.; Park, J.; Song, S. Numerical study of the performance and NOx emission of a diesel-methanol dual-fuel engine using multi-objective Pareto optimization. Energy 2017, 124, 272–283. [Google Scholar] [CrossRef]
  16. Zhang, Z.; Tian, J.; Li, J.; Ji, H.; Tan, D.; Luo, J.; Jiang, Y.; Yang, D.; Cui, S. Effects of Different Mixture Ratios of Methanol-Diesel on the Performance Enhancement and Emission Reduction for a Diesel Engine. Processes 2021, 9, 1366. [Google Scholar] [CrossRef]
  17. Ning, L.; Duan, Q.; Kou, H.; Zeng, K. Parametric study on effects of methanol injection timing and methanol substitution percentage on combustion and emissions of methanol/diesel dual-fuel direct injection engine at full load. Fuel 2020, 279, 118424. [Google Scholar] [CrossRef]
  18. Yilmaz, N.; Sanchez, T. Analysis of operating a diesel engine on biodiesel-ethanol and biodiesel-methanol blends. Energy 2012, 46, 126–129. [Google Scholar] [CrossRef]
  19. Ning, L.; Duan, Q.; Zhanming, C.; Kou, H.; Liu, B.; Yang, B.; Zeng, K. A comparative study on the combustion and emissions of a non-road common rail diesel engine fueled with primary alcohol fuels (methanol, ethanol, and n-butanol)/diesel dual fuel. Fuel 2020, 266, 117034. [Google Scholar] [CrossRef]
  20. Jin, C.; Yao, M.; Liu, H.; Lee, C.-f.; Ji, J. Progress in the production and application of n-Butanol as a biofuel. Renew. Sustain. Energy Rev. 2011, 15, 4080–4106. [Google Scholar] [CrossRef]
  21. Nanthagopal, K.; Kishna, R.; Atabani, A.; Al-Muhtaseb, A.H.; Kumar, G.; Ashok, B. A compressive review on the effects of alcohols and nanoparticles as an oxygenated enhancer in compression ignition engine. Energy Convers. Manag. 2019, 203, 112244. [Google Scholar] [CrossRef]
  22. Szabados, G.; Lukács, K.; Bereczky, A. Investigating Combustion Process of N-Butanol-Diesel Blends in a Diesel Engine with Variable Compression Ratio. Clean Technol. 2021, 3, 618–628. [Google Scholar] [CrossRef]
  23. Satsangi, D.; Tiwari, N. Experimental investigation on combustion, noise, vibrations, performance and emissions characteristics of diesel/n-butanol blends driven genset engine. Fuel 2018, 221, 44–60. [Google Scholar] [CrossRef]
  24. Rakopoulos, D.; Rakopoulos, C.; Giakoumis, E.; Dimaratos, A.; Kyritsis, D. 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]
  25. Ashok, B.; Ashok, S.; Kumar, C. LPG diesel dual fuel engine—A critical review. Alex. Eng. J. 2015, 36, 105–126. [Google Scholar] [CrossRef] [Green Version]
  26. Mobasheri, R.; Peng, Z.; Mirsalim, S. Analysis the effect of advanced injection strategies on engine performance and pollutant emissions in a heavy duty DI-diesel engine by CFD modeling. Int. J. Heat Fluid Flow 2011, 33, 59–69. [Google Scholar] [CrossRef]
  27. Lu, Y.; Liu, Y. Effects of multiple injections on combustion and emissions in a heavy-duty diesel engine at high load and low speed. Adv. Mech. Eng. 2020, 12, 1–15. [Google Scholar] [CrossRef]
  28. 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]
  29. Liu, J.; Yao, A.; Yao, C. Effects of injection timing on performance and emissions of a HD diesel engine with DMCC. Fuel 2014, 134, 107–113. [Google Scholar] [CrossRef]
  30. Liu, B.; Cheng, X.; Liu, J.; Pu, H. Investigation into particle emission characteristics of partially premixed combustion fueled with high n-butanol-diesel ratio blends. Fuel 2018, 223, 1–11. [Google Scholar] [CrossRef]
  31. Bhowmick, P.; Jeevanantham, A.K.; Ashok, B.; Nanthagopal, K.; Perumal, D.D.A.; Viswanathan, K.; Vora, K.C.; Jain, A. Effect of Fuel Injection Strategies and EGR on Biodiesel Blend in a CRDI engine. Energy 2019, 181, 1094–1113. [Google Scholar] [CrossRef]
  32. Paykani, A.; Kakaee, A.; Rahnama, P.; Reitz, R. Effects of diesel injection strategy on natural gas/diesel reactivity controlled compression ignition combustion. Energy 2015, 90, 814–826. [Google Scholar] [CrossRef]
  33. Jiang, X.; Wei, H.; Zhou, L.; Chen, R. Numerical Study on the Effects of Multiple-Injection Coupled with EGR on Combustion and NOx Emissions in a Marine Diesel Engine. Energy Procedia 2019, 158, 4429–4434. [Google Scholar] [CrossRef]
  34. Liu, T.; E, J.; Yang, W.M.; Deng, Y.; An, H.; Zhang, Z.; Pham, M. Investigation on the applicability for adjusting reaction rates of the optimized biodiesel skeletal mechanism. Energy 2018, 150, 1031–1038. [Google Scholar] [CrossRef]
  35. Zhang, Z.; Li, J.; Tian, J.; Xie, G.; Tan, D.; Qin, B.; Huang, Y.; Cui, S. Effects of Different Diesel-Ethanol Dual Fuel Ratio on Performance and Emission Characteristics of Diesel Engine. Processes 2021, 9, 1135. [Google Scholar] [CrossRef]
  36. Gong, C.; Liu, J.; Peng, L.; Liu, F. Numerical study of effect of injection and ignition timings on combustion and unregulated emissions of DISI methanol engine during cold start. Renew. Energy 2017, 112, 457–465. [Google Scholar] [CrossRef]
  37. Zang, R.; Yao, C. Numerical Study of Combustion and Emission Characteristics of a Diesel/Methanol Dual Fuel (DMDF) Engine. Energy Fuel 2015, 29, 3963–3971. [Google Scholar] [CrossRef]
  38. Wang, C.H.; Zhao, D.; Schluter, J.; Holzäpfel, F.; Stephan, A. LES study on the shape effect of ground obstacles on wake vortex dissipation. Aerosp. Sci. Technol. 2017, 63, 245–258. [Google Scholar] [CrossRef] [Green Version]
  39. Han, Z.; Reitz, R. Turbulence Modeling of Internal Combustion Engines Using RNG κ-ϵ Models. Combust. Sci. Technol. 1995, 106, 267–295. [Google Scholar] [CrossRef]
  40. Yakhot, V.; Orszag, S.A.; Thangam, S.; Gatski, T.; Speziale, C. Development of Turbulence Models for Shear Flows by a Double Expansion technique. Phys. Fluids A. 1992, 4, 1510–1520. [Google Scholar] [CrossRef] [Green Version]
  41. Turns, S.R. An Introduction to Combustion; McGraw-Hill, Inc.: New York, NY, USA, 1996. [Google Scholar]
  42. Beale, J.; Reitz, R. Modeling spray atomization with the Kelvin-Helmholtz/Rayleigh-Taylor hybrid model. At. Spray 1999, 9, 623–650. [Google Scholar]
  43. Gonzalez, D.M.; Lian, Z.; Reitz, R. Modeling Diesel Engine Spray Vaporization and Combustion. SAE Trans. 1992, 101, 1064–1076. [Google Scholar]
  44. Naber, J.; Reitz, R. Modeling Engine Spray/Wall Impingement. SAE Trans. 1988, 97, 118–140. [Google Scholar]
  45. Amsden, A.; Orourke, P.; Butler, T. KIVA-2: A Computer Program for Chemically Reactive Flows with Sprays; NASA STI/Recon Technical Report N; Report of Los Alamos National Laboratory: Los Alamos, NM, USA, 1989; Volume 89, p. 27975.
  46. Fan, L.; Cheng, F.; Zhang, T.; Liu, G.; Yuan, J.; Mao, P. Visible-light photoredox-promoted desilylative allylation of a-silylamines: An efficient route to synthesis of homoallylic amines. Tetrahedron Lett. 2021, 81, 153357. [Google Scholar] [CrossRef]
  47. Hiroyasu, H.; Kadota, T. Models for Combustion and Formation of Nitric Oxide and Soot in Direct Injection Diesel Engines. SAE Trans. 1976, 85, 513–526. [Google Scholar] [CrossRef]
  48. Yousefi, A.; Guo, H.; Birouk, M. Effect of swirl ratio on NG/diesel dual-fuel combustion at low to high engine load conditions. Appl. Energy 2018, 229, 375–388. [Google Scholar] [CrossRef]
  49. Liu, H.; Li, J.; Wang, J.; Wu, C.; Liu, B.; Dong, J.; Liu, T.; Ye, Y.; Wang, H.; Yao, M. Effects of injection strategies on low-speed marine engines using the dual fuel of high-pressure direct-injection natural gas and diesel. Energy Sci. Eng. 2019, 7, 375–388. [Google Scholar] [CrossRef] [Green Version]
  50. Zheng, Z.; Li, C.; Liu, H.; Zhang, Y.; Zhong, X.; Yao, M. Experimental study on diesel conventional and low temperature combustion by fueling four isomers of butanol. Fuel 2015, 141, 109–119. [Google Scholar] [CrossRef]
  51. Pan, M.; Huang, R.; Liao, J.; Jia, C.; Zhou, X.; Huang, H.; Huang, X. Experimental study of the spray, combustion, and emission performance of a diesel engine with high n-pentanol blending ratios. Energy Convers. Manag. 2019, 194, 1–10. [Google Scholar] [CrossRef]
  52. Jin, C.; Zhang, X.; Geng, Z.; Pang, X.; Wang, X.; Ji, J.; Wang, G.; Liu, H. Effects of various co-solvents on the solubility between blends of soybean oil with either methanol or ethanol. Fuel 2019, 244, 461–471. [Google Scholar] [CrossRef]
  53. Nayak, S.; Mishra, P.; Noor, M.M. Simultaneous reduction of nitric oxide and smoke opacity in TDI dual fuel engine fuelled with calophyllum-diesel blends and waste wood chip gas for modified inlet valve and injector nozzle geometry. Energy 2019, 189, 116238. [Google Scholar] [CrossRef]
  54. Huang, H.; Zhu, Z.; Zhu, J.; Lv, D.; Pan, Y.; Wei, H.; Teng, W. Experimental and numerical study of pre-injection effects on diesel-n-butanol blends combustion. Appl. Energy 2019, 249, 377–391. [Google Scholar] [CrossRef]
  55. Li, W.; Ji, J.; Huang, L.; Guo, Z. Global dynamics of a controlled discontinuous diffusive SIR epidemic system. Appl. Math. Lett. 2021, 121, 107420. [Google Scholar] [CrossRef]
  56. Zuo, H.; Tan, J.; Wei, K.; Huang, Z.; Zhong, D.; Xie, F. Effects of different poses and wind speeds on wind-induced vibration characteristics of a dish solar concentrator system. Renew. Energy 2021, 168, 1308–1326. [Google Scholar] [CrossRef]
  57. Zuo, H.; Liu, G.; E, J.; Zuo, W.; Wei, K.; Hu, W.; Tan, J.; Zhong, D. Catastrophic analysis on the stability of a large dish solar thermal power generation system with wind-induced vibration. Sol. Energy 2019, 183, 40–49. [Google Scholar] [CrossRef]
  58. Zuo, H.; Zhang, B.; Huang, Z.; Wei, K.; Tan, J. Effect analysis on SOC values of the power lithium manganate battery during discharging process and its intelligent estimation. Energy 2022, 238, 121854. [Google Scholar] [CrossRef]
  59. Ma, Y.; Liu, C.; E, J.; Mao, X.; Yu, Z. Research on modeling and parameter sensitivity of flow and heat transfer process in typical rectangular microchannels: From a data-driven perspective. Int. J. Therm. Sci. 2022, 172, 107356. [Google Scholar] [CrossRef]
  60. E, J.; Cai, L.; Li, J.; Ding, J.; Chen, J.; Luo, B. Effects analysis on the catalytic combustion and heat transfer performance enhancement of a non-premixed hydrogen/air micro combustor. Fuel 2022, 309, 122125. [Google Scholar] [CrossRef]
  61. Zhang, B.; Zuo, H.; Huang, Z.; Tan, J.; Zuo, Q. Endpoint forecast of different diesel-biodiesel soot filtration process in diesel particulate filters considering ash deposition. Fuel 2020, 272, 117678. [Google Scholar] [CrossRef]
  62. E, J.; Zhao, M.; Zuo, Q.; Zhang, B.; Zhang, Z.; Peng, Q.; Han, D.; Zhao, X.; Deng, Y. Effects analysis on diesel soot continuous regeneration performance of a rotary microwave-assisted regeneration diesel particulate filter. Fuel 2020, 260, 116353. [Google Scholar] [CrossRef]
  63. Feng, C.; Deng, Y.; Chen, L.; Han, W.; E, J.; Wei, K.; Han, D.; Zhang, B. Hydrocarbon emission control of a hydrocarbon adsorber and converter under cold start of the gasoline engine. Energy 2022, 239, 122138. [Google Scholar] [CrossRef]
  64. Chen, L.; Deng, Y.; Feng, C.; Han, W.; E, J.; Wang, C.; Han, D.; Zhang, B. Effects of zeolite molecular sieve on the hydrocarbon adsorbent performance of gasoline engine of during cold start. Fuel 2022, 310, 122427. [Google Scholar] [CrossRef]
  65. E, J.; Luo, J.; Han, D.; Tan, Y.; Feng, C.; Deng, Y. Effects of different catalysts on light-off temperature of volatile organic components in the rotary diesel particulate filter during the regeneration. Fuel 2022, 310, 122451. [Google Scholar] [CrossRef]
  66. Zhang, Z.; Ye, J.; Lv, J.; Xu, W.; Tan, D.; Jiang, F.; Huang, H. Investigation on the effects of non-uniform porosity catalyst on SCR characteristic based on the field synergy analysis. J. Environ. Chem. Eng. 2022, 10(2), 107056. [Google Scholar] [CrossRef]
  67. E, J.; Liu, G.; Zhang, Z.; Han, D.; Chen, J.; Wei, K.; Gong, J.; Yin, Z. Effect analysis on cold starting performance enhancement of a diesel engine fueled with biodiesel fuel based on an improved thermodynamic model. Appl. Energy 2019, 243, 321–335. [Google Scholar] [CrossRef]
  68. Zhang, Z.; E, J.; Deng, Y.; Pham, M.; Zuo, W.; Peng, Q.; Yin, Z. Effects of fatty acid methyl esters proportion on combustion and emission characteristics of a biodiesel fueled marine diesel engine. Energy Convers. Manag. 2018, 159, 244–253. [Google Scholar] [CrossRef]
Figure 1. The geometric model of the combustion chamber in the CONVERGE environment.
Figure 1. The geometric model of the combustion chamber in the CONVERGE environment.
Processes 10 00060 g001
Figure 2. Comparison of the in-cylinder pressure curves of different basic grids.
Figure 2. Comparison of the in-cylinder pressure curves of different basic grids.
Processes 10 00060 g002
Figure 3. Schematic diagram of engine test.
Figure 3. Schematic diagram of engine test.
Processes 10 00060 g003
Figure 4. Engine validation results for cylinder pressure and heat release rate.
Figure 4. Engine validation results for cylinder pressure and heat release rate.
Processes 10 00060 g004
Figure 5. Engine validation results for soot and NOx emissions.
Figure 5. Engine validation results for soot and NOx emissions.
Processes 10 00060 g005
Figure 6. The layout of the injector location in the CONVERGE environment.
Figure 6. The layout of the injector location in the CONVERGE environment.
Processes 10 00060 g006
Figure 7. Comparisons of cylinder pressure under different load conditions.
Figure 7. Comparisons of cylinder pressure under different load conditions.
Processes 10 00060 g007
Figure 8. Comparison of HRR under different load conditions.
Figure 8. Comparison of HRR under different load conditions.
Processes 10 00060 g008
Figure 9. Comparisons of cylinder temperature changes under different load conditions.
Figure 9. Comparisons of cylinder temperature changes under different load conditions.
Processes 10 00060 g009
Figure 10. Comparisons of the cylinder temperature distribution field under different load conditions.
Figure 10. Comparisons of the cylinder temperature distribution field under different load conditions.
Processes 10 00060 g010
Figure 11. Comparisons of soot emission under different load conditions.
Figure 11. Comparisons of soot emission under different load conditions.
Processes 10 00060 g011
Figure 12. Comparisons of NOx emission under different load conditions.
Figure 12. Comparisons of NOx emission under different load conditions.
Processes 10 00060 g012
Figure 13. Comparison of NOx distribution field changes under different load conditions.
Figure 13. Comparison of NOx distribution field changes under different load conditions.
Processes 10 00060 g013
Figure 14. Comparison of HC changes under different load conditions.
Figure 14. Comparison of HC changes under different load conditions.
Processes 10 00060 g014
Figure 15. Comparison of the CO percentage change under different load conditions.
Figure 15. Comparison of the CO percentage change under different load conditions.
Processes 10 00060 g015
Table 1. Engine specifications and boundary conditions.
Table 1. Engine specifications and boundary conditions.
TypeValueTypeValue
Bore × stroke (mm)190 × 210Compression ratio14
Number of cylinders4Head temperature (K)553
Engine speed (rpm)2000Piston temperature (K)423
Effective power (kW)220Wall temperature (K)433
Nozzle radius (mm)0.26Temperature at IVC (K)341
Fuel injection holes8Pressure at IVC (bar)1.97
Cylinder diameter (mm)190Turbulent kinetic energy (m2/s2)62.0271
Connecting rod (mm)410Turbulent dissipation (m2/s3)17183.4
Table 2. Fuel physical properties.
Table 2. Fuel physical properties.
Performance IndexDieselMethanoln-Butanol
Latent heat of gasification (kJ/kg)2601162.2919.6
Density (kg/m3) at 20 °C835792810
Auto-ignition temperature (°C)250463343
Low calorific value (MJ/kg)42.520.130.63
Cetane number513.817
Oxygen content (%) by weight05021.62
Stoichiometric air/fuel ratio14.36.511.21
Kinematic viscosity (40 °C) (mm2/s)2.720.582.22
Table 3. Specifications of the main measuring instruments.
Table 3. Specifications of the main measuring instruments.
ItemContentAccuracyUncertainty (%)
Electric dynamometerNIDY S22-2/0525-1BV-1Torque: ±0.5% F.S; Speed: ±1 r/min±0.2
Dynamometer control systemPUMA OPEN1.4.1±0.5% F.S±0.2
Air flowmeterTOCEIL 20N125±1%±0.1
Diesel flowmeterTOCEIL CMFG0100.12%±0.2
Temperature sensorThermojunction type±0.5 °C±0.1
Pressure sensorPiezoresistance type±0.5% F.S±0.5
Emissions analyzerAVL AMAi60±1.0% F.S±0.2
Combustion analyzerDEWE-2010CA/±0.2
Injection measuring instrumentEFS-IFR600±0.5%±0.5
Table 4. Uncertainty of measured parameters.
Table 4. Uncertainty of measured parameters.
MeasurementsMeasuring RangeAccuracyUncertainty (%)
Engine speed1–2000 rpm±0.2%±0.24
Pressure sensor0–25 MPa±10 kPa±0.5
Exhaust gas temperature0–1000 °C±1 °C±0.25
HC emission0–20,000 ppm±10 ppm±0.11
NOx emission0–5000 ppm±10 ppm±0.53
CO emission0–10% vol±0.03%±0.32
Soot emission0–9 FSN±0.1 FSN±2.8
Air flow rate0–33.3 kg/min±1%±0.5
Fuel flow rate0.5–100 L/h±0.04 L/h±0.5
Table 5. Parameters of the pre-injection strategy.
Table 5. Parameters of the pre-injection strategy.
CaseLoad (%)SOPI
(deg ATDC)
MpreSOMI
(deg ATDC)
Mmain
150--−91.0
250−15~−450.5−90.5
350−450.1~0.9−90.9~0.1
4100--−91.0
5100−15~−450.5−90.5
6100−450.1~0.9−90.9~0.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, Z.; Li, L. Effect of Pre-Injection on Combustion and Emission Characteristics of a Diesel Engine Fueled with Diesel/Methanol/n-Butanol Blended Fuel. Processes 2022, 10, 60. https://doi.org/10.3390/pr10010060

AMA Style

Wang Z, Li L. Effect of Pre-Injection on Combustion and Emission Characteristics of a Diesel Engine Fueled with Diesel/Methanol/n-Butanol Blended Fuel. Processes. 2022; 10(1):60. https://doi.org/10.3390/pr10010060

Chicago/Turabian Style

Wang, Zhiqiang, and Lijun Li. 2022. "Effect of Pre-Injection on Combustion and Emission Characteristics of a Diesel Engine Fueled with Diesel/Methanol/n-Butanol Blended Fuel" Processes 10, no. 1: 60. https://doi.org/10.3390/pr10010060

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