Next Article in Journal
Tourism Enterprises Marketing Management and Upgrading Situation of Tourist Sites to Achieve Sustainable Regional Economic Development
Previous Article in Journal
Population Shrinkage, Public Service Levels, and Heterogeneity in Resource-Based Cities: Case Study of 112 Cities in China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Diesel Spray: Development of Spray in Diesel Engine

by
Djati Wibowo Djamari
1,
Muhammad Idris
2,
Permana Andi Paristiawan
3,
Muhammad Mujtaba Abbas
4,
Olusegun David Samuel
5,6,*,
Manzoore Elahi M. Soudagar
7,8,
Safarudin Gazali Herawan
9,
Davannendran Chandran
10,
Abdulfatah Abdu Yusuf
11,
Hitesh Panchal
12 and
Ibham Veza
10,*
1
Mechanical Engineering Study Program, Sampoerna University, Jakarta 12780, Indonesia
2
PT Perusahaan Listrik Negara (Persero), Engineering and Technology Division, Jakarta 11420, Indonesia
3
Research Center for Metallurgy, National Research and Innovation Agency, South Tangerang 15314, Indonesia
4
Department of Mechanical Engineering, University of Engineering and Technology (New Campus), Lahore 54890, Pakistan
5
Department of Mechanical Engineering, Federal University of Petroleum Resources, P.M.B 1221, Effurun 330102, Nigeria
6
Department of Mechanical Engineering, University of South Africa, Science Campus, Private Bag X6, Florida 1709, South Africa
7
Department of Mechanical Engineering and University Centre for Research & Development, Chandigarh University, Mohali 140413, India
8
Department of Mechanical Engineering, School of Technology, Glocal University, Delhi-Yamunotri Marg, SH-57, Mirzapur Pole, Saharanpur 247121, India
9
Industrial Engineering Department, Faculty of Engineering, Bina Nusantara University, Jakarta 11480, Indonesia
10
Mechanical Engineering, Universiti Teknologi PETRONAS, Seri Iskandar 32610, Malaysia
11
Department of Mechanical Engineering, University of Liberia, Monrovia 1000, Liberia
12
Department of Mechanical Engineering, Government Engineering College, Patan 384265, India
*
Authors to whom correspondence should be addressed.
Sustainability 2022, 14(23), 15902; https://doi.org/10.3390/su142315902
Submission received: 23 October 2022 / Revised: 20 November 2022 / Accepted: 28 November 2022 / Published: 29 November 2022
(This article belongs to the Topic Clean and Low Carbon Energy)

Abstract

:
Research and development in the internal combustion engine (ICE) has been growing progressively. Issues such as air pollution, fuel cost, and market competitiveness have driven the automotive industry to develop and manufacture automobiles that meet new regulation and customers’ needs. The diesel engine has some advantages over the gasoline or spark ignition engine, including higher engine efficiency, greater power output, as well as reliability. Since the early stage of the diesel engine’s development phase, the quest to obtain better atomization, proper fuel supply, and accurate timing control, have triggered numerous innovations. In the last two decades, owing to the development of optical technology, the visualization of spray atomization has been made possible using visual diagnostics techniques. This advancement has greatly improved research in spray evolution. Yet, a more comprehensive understanding related to these aspects has not yet been agreed upon. Diesel spray, in particular, is considered a complicated phenomenon to observe because of its high-speed, high pressure, as well as its high temperature working condition. Nevertheless, several mechanisms have been successfully explained using fundamental studies, providing several suggestions in the area, such as liquid atomization and two-phase spray flow. There are still many aspects that have not yet been agreed upon. This paper comprehensively reviews the current status of theoretical diesel spray and modelling, including some important numerical and experimental aspects.

1. Introduction

Energy transition and environmental issues on the restrictions of emissions have been paid much more attention in most countries across the world [1,2,3,4]. CO2 emissions have been addressed as the dominant pollutant that contribute to climate change [5]. Hence, engine manufacturers should improve the performance of internal combustion engines by developing efficient engines with lower emissions [6,7]. Regardless of the stringent emissions regulation imposed by several countries, diesel fuel remains the major source of energy for heavy duty engines. The basic explanation for this is because diesel engines have a relatively higher thermal efficiency spark ignition (SI) than gasoline engines, which is especially important for heavy duty purposes.
In the diesel engine, mixture formation between air and fuel plays a vital role in the process of combustion, thus affecting emission qualities [8]. To enhance the combustion and lower the emission, it is essential to have comprehensive knowledge of the mixture formation. Prior to acquiring this knowledge, a thorough understanding of spray atomization must be first understood. Studies on diesel spray remains a crucial part of research in the internal combustion engine. The quality of the ignition is substantially influenced by local and temporal processes, as well as the completeness of combustion. The pollutants in the cylinder interact and break down, generating the components in the exhaust gas. This is specifically true for carbon monoxide (CO) and unburned hydrocarbon (UHC) emissions. The process of fuel injection controls how well the engine performs in terms of combustion and emissions. Note that the injection system from the high-pressure common-rail diesel engine is currently viewed as one of the most viable ways to satisfy the increasingly strict emission standards for diesel engines. This illustrates the importance of diesel spray in the development of the diesel engine.
Spray is a process in which liquid is forced out of a particular container so that it becomes a mass of small-liquid drops. In the direct-injection internal combustion engine, fuel, which is still in its liquid phase, is injected by the injector into a gaseous condition in the cylinder. In a compression ignition engine, such a phenomenon is commonly known as diesel spray. The injected liquid fuel that is now inside the combustion chamber then begins to interact with the in-cylinder air instantaneously, leading to the dispersion of the liquid phase. As this liquid phase vaporizes, a flammable mixture is then generated to produce great power inside the chamber. The whole process is known as the spray evolution process, consisting of liquid-fuel breakup, atomization, and evaporation. A detailed knowledge in spray evolution is important as the mixture preparation, which later determines the quality of the combustion and emission formation, depends greatly on this process.

2. Diesel Combustion Process

Unlike gasoline engines, that use a spark plug to initiate the combustion, diesel engines work based on the principle of autoignition, as shown in Figure 1. Following fuel in the liquid form being injected into the cylinder, this fuel then mixes with the in-cylinder air at a high temperature and pressure, achieved by the high volumetric compression of the air. The typical compression ratios in diesel engines ranges from 15 to 22. Figure 1 shows one complete working cycle of a typical diesel engine via four strokes. The crankshaft must complete two revolutions (720°) per cycle of operation.
Diesel engines can operate in high compression ratios, while their counterparts, gasoline engines, operate only in low to medium ratios to avoid engine knock. Due to these relatively higher compression ratios, diesel engines are more fuel efficient, making them a natural choice for heavy-duty purposes. In modern diesel engines that use common-rail technology, a high-pressure injector is used to insert high-velocity liquid fuel into the hot compressed air. The fuel pressure is raised to the desired level using a pump, which moves the liquid fuel from a tank to multiple injectors. As the engine load is subject to variation in the engine speed while the vehicles are operated, the load thus needs to be adjusted by regulating the fuel amount to be injected. Each injector then spreads the liquid fuel in small droplets. As the liquid fuel penetrates through the combustion chamber, this fuel then breaks down to smaller droplets, evaporates, as well as mixes with the in-cylinder gas. This is when the autoignition starts to happen, before the main combustion finally takes place in the chamber.
As mentioned above, the increase in injection pressure is considered as a promising approach to enhance combustion behaviors [9,10,11,12,13]. High-pressure injected diesel fuel could improve the engine performance and reduce particulate matter [14,15,16,17,18]. In addition to that, the increased injection pressure tends to accelerate the vaporization of droplets and the preparation of the combustible mixture [19,20,21,22,23]. As expected, this would reduce the ignition delay. Note that the process of combustion begins with the autoignition of the fuel and air mixture. The characteristics of autoignition are affected by the chemical as well as the physical properties of the fuel. As it is more difficult to observe those two properties macroscopically, the ignition delay is hence used as a key measure of the autoignition. However, in typical combustion, autoignition occurs in two stages that are categorized by these properties: physical and chemical.
The first stage of autoignition is determined by the physical properties of the liquid spray. The delay between the injection start, and the combustible mixture formation, is known as the physical delay. Such a delay is controlled by its physical properties (density, surface tension, and specific heat). Furthermore, the structures of the spray flow in the combustion chamber also have substantial effects on the physical delay. Take small- and large-scale hydrodynamic structures, for example, these types of structures tend to drive the entrainment of the air into the fuel jet in a turbulent environment. The liquid jet starts to break down and the spray begins to evaporate, caused by aerodynamic interactions and ambient air interactions, respectively.
The second stage of the autoignition is determined by the fuel chemical properties. The delay between the mixture formation and its ignition is known as the chemical delay. Such a delay is controlled by its chemical properties i.e., the molecular structure, adiabatic flame temperature, C/H/O ratio, and the sooting tendency. Therefore, the total ignition delay is the combination between physical and chemical delays. As far as the chemical kinetics are concerned, the process of chemical kinetics in the diesel engine is a complicated mechanism as it involves many simultaneous reactions. Although the rate of heat release that causes a rapid rise in pressure and temperature leading to autoignition is easy to observe, the chemical mechanisms involved microscopically are complex. These mechanisms include initiation, propagation, and the termination processes of the chain.
Firstly, the chain begins its initiation reaction by building a group of radical species from the reactants. The radicals then react with each other, as well as with other stable species, to form more radicals, resulting in the propagation of the chain reactions. Of these chain propagation reactions, chain branching reactions are of great importance as they determine the self-propagating flame that leads to combustion. These chain branching reactions can have an explosive effect on a combustion system by producing radicals that control the overall reaction rate. Lastly, the stable products are formed afterwards, which indicate the termination process of the reactions. The overall chain reactions thus consist of a complicated process of simultaneous chemical reactions, which are also subject to the mixture physical state.
In diesel or compression ignition engines, the combustion occurs in two phases following the autoignition of the mixture inside the cylinder [24,25]. They are premixed-controlled and mixing-controlled combustion. The premixed-controlled combustion occurs in a fuel-rich region and produces partially oxidized fuel fragments that diffuse outwardly, leading to the mixing-controlled diffusion flame that surround the premixed core region. The first stage (heat release along with temperature rise) takes place in a fuel-rich premixed region. The overall HRR is primarily controlled by the fuel mixing rate and oxidizer. Typical direct injection in the diesel engine is schematically illustrated in Figure 2.
As the liquid fuel jet penetrates the cylinder, it entrains the surrounding gas. The liquid spray is then heated up and evaporation occurs at the spray edge. The more air enters the spray, the higher the evaporation rate of the spray. The combustion then starts, while the mixture reaches a particular equivalence ratio. In the region of the diffusion flame, the orange color represents the highest temperature, which is believed to be the primary location of NO production [27]. This region also oxidizes a great amount of soot that is generated in the fuel-rich region. This diffusion layer flame can be identified using the OH* chemiluminescence technique for laser-based optical combustion diagnostics [28,29,30].
There are a number of strategies to improve the diesel combustion process. The introduction of pilot fuel is believed to be able to reduce the combustion noise as it shortens the premixing process [31]. The pilot fuel could reduce the pressure-rise rate by allowing only a small amount of liquid fuel to be introduced for the premixed-combustion mixture, leading to lower NOx emission [32,33]. Just after this pilot fuel has been introduced, the appropriate amount of fuel should be injected immediately to avoid the formation of soot. A post injection is then needed, avoiding the formation of soot [34,35,36]. In addition to that strategy, the multiple injections approach was also proposed to obtain the desired MFR shape. The multiple injections strategy allows the injector to have a shorter injection duration for each split injection, allowing fuel to blend sufficiently with the air so that an improved mixture can be achieved. Such a strategy also increases the IMEP, which was found to reduce the emissions. The multiple injections strategy can also prevent the impingement of the spray onto the wall chamber, resulting in low hydrocarbon (HC) pollutant. The multiple injections concept in the diesel engine has nowadays been used widely, due to its several advantages. Its superiority compared to the single injection mode is achieved through three different stages: (i) pre-injection allows a small quantity of fuel to be injected, leading to a reduction in combustion noise and NOx emissions; (ii) the main injection provides optimized torque release without producing excessive particles from the combustion in the chamber; and (iii) post-injection gives a better soot oxidation rate by letting other small fuel mass to be injected after the main injection period.

3. Spray Atomization

Diesel spray is a crucial element of compression ignition engine operation, particularly during the injection of fuel. Many studies have been conducted focusing on spray atomization [37,38,39,40,41]. Several factors are known to be responsible for spray atomization. They were: (i) the aerodynamic forces; (ii) internal turbulent; (iii) boundary mutation at the exit of the nozzle; and (iv) internal flow cavitation. With the help of the optically-accessible-measurement technique, none of the factors mentioned above were found to influence the spray atomization independently. In fact, it is believed that a combination of some dominant factors is actually responsible collectively.
An aerodynamic force is defined as the force that a body experiences from the air (or other gas), where it is submerged. This force results from the relative velocity of the body and the gas. Aerodynamic force is produced by two factors: the normal force, which results from pressure acting on the body’s surface; and the shear force, which is a result of the viscosity of the gas, and is also referred to as skin friction. Shear force moves parallel to the surface, while pressure moves perpendicular to the surface. Both factors are regional. The pressure, as well as shear forces summed over the entire exposed region of the body, constitute the body’s net aerodynamic force.
When diesel spray leaves the orifice of the nozzle, aerodynamic-triggered break-up gradually contributes to the production of the mixture. It is important to remember that coherent liquid does not instantly break up into small droplets, and that a zone with ligaments and concentrated big droplets is present close to the nozzle exit. The internal turbulence, along with cavitation, have a more significant impact on the disintegration of the fuel during this process than compared to aerodynamic forces, hence it should be incorporated in the primary break-up process. Aerodynamic force causes the secondary break-up, which is the creation of smaller droplets, when the spray increases the break-up length. The balance between surface tension and aerodynamics is crucial to this process. As previously stated, the aerodynamic force, that primarily serves as ambient gas entrainment, dominates the break-up process, particularly for secondary break up. As a result, researchers pay close attention to the air entrainment.
Cavitation occurs as a result of sudden alterations in the shape of the internal nozzle and pockets from low-static pressure. Vapor phase disruptions, internal flow circumstances, wall roughness, as well as any surface-level microscale manufacturing flaws, all contribute to turbulence. Cavitation would raise the liquid velocity at the exit of the nozzle, which will affect how the spray forms and how it is made. Enhanced spray development is generally believed to result in a more complete combustion, less fuel consumption, and fewer emissions. However, because of its impact on the outgoing jet, cavitation can reduce flow effectiveness (discharge coefficient). Additionally, expanding cavitation bubbles on the inside of the orifice can lead to material degradation, reducing the injector’s lifespan and effectiveness.
Numerous studies have investigated internal flow cavitation [42,43,44]. It is believed that the disintegrations of the needle lift in the cavitation bubbles speed up the spray breakup. The cavitation of the nozzle hole could increase the spray angle and decrease the penetration of the spray tip. Interestingly, the nozzle hole inlet geometry could also influence the intensity of cavitation. It is important to remember that liquid fuel disintegrates into ligaments and large droplets at the stage of internal flow, before the spray comes out of the nozzle. This process, known as primary break-up, is caused by the internal flow cavitation and turbulence instability [45,46,47]. Figure 3 is break-up regimes of droplet.
The break-up mechanisms of these two factors usually occur simultaneously. Turbulent eddies formed the droplets in small aerodynamic condition, such as in nozzle-hole internal flow. It took place while the surface energy was smaller, compared to the radial fluctuation kinetic energy of the eddy flow velocity, which greatly affects the turbulent effect. It is also known that when the flow velocity is high, the cavitation starts to form. The cavitation takes place since the liquid phase vaporizes instantly when the pressure of local static is lower than vapor pressure (the static pressure decreases when the flow velocity increases). Thus, a two-phase flow emerges in the interior of the nozzle hole once the cavitation is formed. This cavitation phenomenon is characterized by the vaporization of low temperature volatile elements.
The spray penetration was time-dependence, being proportional to the evolution timing. Droplets were formed in the spray during the second phase. In this phase, despite the drop in its penetrating velocity, the spray still continued to penetrate further inside the chamber because of energy from the late injection of high momentum fuel. As discussed previously, the break-up (secondary) was primarily initiated by the forces of aerodynamic. It was widely known that aerodynamic forces were greatly dependent on air entrainment. Hence, it is important to investigate the air entrainment in diesel spray. The surrounding air mostly entrained into the spray via the upstream.
The spray and wall interaction in the diesel engine has gained a great deal of attention recently. The impingement of spray on the wall primarily happens in a small-bore engine (high speed) and is believed to have a major effect on mixture formation, combustion, as well as emission. At this impinging condition, the spray atomization undergoes different processes compared to those of free spray. Two main primary differences are the improvement of the secondary break-up process due to smaller droplets formation at a large Weber number, and the reduction of unburned hydrocarbon and soot emission, due to liquid film formation at a low Weber number [49,50,51,52,53]. The interaction of wall and spray, mainly taking place in the small-bore high-speed diesel engine, not only has significant effects on mixture formation, but also on the combustion, as well as the emission process in the cylinder. In order to clarify the effect of spray impingement, an analysis of the two-dimensional piston cavity, along with flat wall impinging spray flame, is usually performed in a high pressure and temperature constant-volume engine-like combustion chamber [54,55,56,57,58]. As a basis for evaluation, free spray flame is normally also conducted.
To recognize the impact of diesel spray on the engine, the most influential factors within the spray structure must be first understood. Those determinant factors are illustrated macroscopically in Figure 4. Near the nozzle tip is the region in which the fuel in the liquid form is dense and uniform [59,60,61]. Below this dense area, fuel then begins to disperse, and waves begin to emerge at the edge of the spray. Lines of liquid also start to exist and the pitches in the middle of these stripes increase additionally downstream. Afterwards, spray and air inside the chamber interact immediately. As a result of the penetrating spray, the air surrounding it then compresses the spray clusters to be compact. Simultaneously, the spray is transferring its momentum to the air leading to further dispersion of the spray. This is called the air entrainment effect [62,63,64,65,66].
  • Spray core: high dense spray region, the closer to the injector is, the denser it is.
  • Spray angle: a bigger spray angle gives wider spatial distribution.
  • Break-up length: a part of the liquid in the spray that does not disintegrate.
  • Spray tip penetration: the macroscopic development of diesel spray.
Large angle spray is favored and is formed by short spray core and break-up length. Long break-up length, on the other hand, is avoided as it causes a narrow unsteady spray [68,69,70,71]. Moreover, due to its longer spray characteristics, the fuel might stick to the inner side of the chamber when it is impinged to the in-cylinder wall, leading to high hydrocarbon and particulate matter emissions. As the formation of the mixture continues, the spray volume is increasing parallel to the increase of an entrained air. As a consequence, the spray penetrates further due to the influences from such air entrainment effect and spray velocity. This phenomenon can be observed and analyzed by optical diagnostic techniques, which will be explained later.

4. Factors Affecting Spray Evolution

Injector technology plays a crucial role in determining the quality of atomization, particularly the injector type and its geometry [40,72,73,74,75]. For diesel engines, two types of injectors are normally used, namely, the piezo and the solenoid injector. The piezo injector has a relatively quicker response and opening, which results in shorter injection delay, thus allowing the injected fuel to blend with the air quicker. As a result, mixture formation can be improved leading to better engine performance and reduced emission pollutants. Furthermore, the quicker response of the piezo injector tends to ease the fuel mass delivery control and to enhance the profile of MFR.
In terms of injector geometry, nozzle configuration is classified into three main categories: the mini and micro sac, as well as the valve covered orifice (VCO) [40,76,77,78,79]. The difference in the injector nozzle geometry greatly affects the cavitation and evolution of the spray. The illustration of those configurations is displayed in the following Figure 5.
The sac refers to the part of an injector that looks like a small bag and contains liquid fuel [51,80,81,82]. As a result, nozzles with this geometry can lead to unburned hydrocarbon left in the sac. Despite its drawback, the nozzles with sac configuration tend to have superior spray quality compared to VCO nozzle geometry, which have no sac configuration [40,76,77,83,84]. This is because the holes in the sac configuration are positioned just below the needle seat, thus minimizing the loss from throttling. Additionally, VCO is also subject to working conditions and can consequently lead to the oscillation of the needle. In more than one injector, for example, the pressure tends not to be uniformly distributed around the needle in the course of the opening. As a result, the holes-cycles mass flow varies. The VCO injector, however, is able to provide a specific fuel amount and govern the time of injection more accurately, since no sac volume has to be filled. Clearly, the sac and VCO configuration has different effects in the spray evolution. In terms of the cavitation, for example, vortex cavitation occurs in the sac configuration, while conventional cavitation takes place in VOC nozzle geometry.
In order to examine the impact of both the injector hole size and the pressure of the injection, Lacoste [85] evaluated the droplet velocity and SMD with the hole diameters set to be 0.1 and 0.2 mm, whereas the injection pressures were varied at two opposite conditions: low pressure (60 MPa), and high pressure (160 MPa). The results revealed that the smaller hole managed to reduce the droplet size and at the same time increase its velocity. The cavitation was also improved, thus leading to the enhanced turbulence effect. Therefore, the better atomization quality of the spray can be achieved by simultaneously using higher injection pressure and reduced hole size. The nozzle hole convergence can significantly affect the spray behavior. Note that the hole geometry could affect the spray behavior. Different nozzle structure could lead to different spray behavior [86,87,88,89,90].
The injection pressure is critical in determining the spray evolution [91,92,93,94,95]. A high-inertia effect from high-injection pressure results in high-velocity spray, thus creating better dispersion by forming relatively smaller droplets [53,96,97,98,99]. Higher-injection pressure could deliver the same amount of fuel relatively faster [100,101,102,103,104]. Moreover, the rise in injection pressure also causes back pressures of the spray that interact with the air [105,106,107,108,109]. This will lead to reduced breakup time and length. As a result, faster propagation and better atomization can be achieved, despite higher fuel concentration at the beginning of the injection. Note that the increased injection pressure can allow the entrainment of the air from the surrounding gas to completely develop, enabling droplets to interact with ambient gas [54,107,110,111,112].
Ambient pressure, along with density, have a great influence on macroscopic attributes of the spray [113,114,115,116,117]. Additionally, the rise in ambient density could reduce the length of spray penetration as well as widen the angle of the spray cone [39,65,114,118,119]. The ambient temperature primarily affects the spray behavior in terms of its evaporation rate, both microscopically and macroscopically. Hot ambient air inside the cylinder could improve the evaporation of the spray outer periphery and consequently decreasing the spray cone angle [39,107,120,121,122]. Increasing ambient temperature could lead to enhanced evaporation rate and thus considerably shorten the spray penetration length [123,124,125,126,127].
During the injection start, where the fuel in the liquid form is relatively dense, the spray penetration rate is not enhanced by higher ambient temperature; instead, it is influenced by the gas density [114,128,129,130,131]. As the spray penetrates downstream, the liquid fuel becomes less dense, thus increasing both heat as well as momentum transfers by the side of the front periphery. Although lower gas density could raise the penetration rate, the droplet clusters at the front periphery are separated more quickly, resulting in relatively far shorter spray length penetration. The raise in evaporation rate under hot ambient temperature could reduce the SMD. Note that higher ambient air temperature could reduce droplets size at the periphery. The center of the droplet, however, is not influenced by raised temperature owing to the high density.
Higher fuel temperature tends to produce vapor bubble and this phenomenon is known as cavitation [125,132,133]. In the liquid fuel injection behavior, cavitation combined with turbulence can improve spray performance in terms of the liquid atomization, droplets breakup and air entrainment. Under a particular case, such as an extreme condition, cavitation could result in a hydraulic flip that can worsen atomization.
One fuel property that significantly affects the combustion inside the cylinder is viscosity [134,135,136,137,138]. It influences the mixture formation, MFR, and particle size distribution. High viscosity fuel tends to stabilize the spray and delays the fuel dispersion as extra energy is needed to offset the effect of viscous force [139,140,141,142,143]. In contrast, low viscosity fuel has better dispersion resulting in shorter penetration length but wider cone angle [114,130,144,145,146]. Biodiesel, for instance, could hinder the fuel atomization owing to its high viscosity as well as surface tension content [147,148,149,150,151].
Surface tension is another determinant factor that influences the atomization as it signifies the energy amount needed to increase bubbles [152,153,154,155]. While surface tension affects the fuel surface area, volatility influences the spray evolution and mixture formation [114,156,157,158,159]. High volatile fuel tends to increase the evaporation rate and lowers SMD thus reducing the spray penetration length [160,161,162,163,164,165]. Low volatile fuel, on the other hand, has relatively lower evaporation rate, larger droplets size, thus increasing spray penetration length, which in some cases could lead to the wall impingement [166,167,168,169,170].

5. Modeling of Turbulent Diesel Spray and Combustion

The turbulence behavior of diesel spray and the combustion process play a considerable part in influencing the engine performance, combustion, as well as emission formation [171,172,173,174]. The physical understanding of these two concepts (diesel spray and combustion) brings a substantial contribution to the modeling in diesel engine combustion. The success of computational models is greatly dependent on how accurate the models can represent the actual combustion process inside the cylinder. Computational fluid dynamic (CFD) models have been proposed, improved, and implemented to capture the physical processes that determine turbulent combustion in diesel engines.
For the atmospheric pressure turbulent flames modeling, until today, most of the turbulent flame models have been approached using simple geometric configurations towards laboratory-scale, in which the flames pressure is assumed to be near atmospheric. This is because in such a condition the thermochemical conditions are well known, and sufficient measurement data are available. To investigate piloted flames, Sandia has generously provided turbulent flame data to facilitate several modeling studies in engine combustion. These data include Sandia C, D, E, and F flames, that are non-luminous, turbulent, and non-premixed. The fuel jet, along with pilot velocities, rise from flame C to F, and so do their local extinction effects. Therefore, while flame D shows mild extinction effects, flame F shows near global extinction effects. The flames provided by Sandia are suitable for modeling studies of turbulence chemistry interactions (TCI). The Damköhler number is often used to characterize the importance of TCI. The number represents the ratio of hydrodynamic to chemical time scales. Low-to-intermediate Damköhler numbers are of particular concern here as they indicate the important phenomena occurs in turbulent and chemistry reactions.
Dispersion, collisions of dispersed phase inter-particle, the modification of continuous phase turbulence, evaporation, mixing, as well as combustion, all take place concurrently in a true turbulent spray flame. It is a huge modelling challenge to deal with all of these intricacies and their relationships. Therefore, it is sensible to aim for advancement in individual sub-areas like breakdown, dispersion, mixing, and combustion, which cannot be completely viewed in isolation. Additionally, the benefits and drawbacks of the general modelling approach must be taken into consideration, including probability density function (PDF), direct numerical simulation (DNS), and large eddy simulation based on Reynolds averaged equations [175].
Several modeling studies have been proposed to improve the PDF-based models by exploring variants of transported modeling, such as composition and velocity-composition-joint PDF, advanced mixing models, improved and faster numerical algorithms, parallelization strategies, radiation modeling, and detailed chemistry. The turbulence modeling and flow structures configurations have been approached with both Reynolds average simulation (RAS) and large eddy simulation (LES). While RAS uses the PDF model, LES uses FDF, which stands for filtered density function. The multi-environment Eulerian-field PDF method, or often shortened MEPDF, has been developed as an alternative to particle-based Monte Carlo methods. The MEPDF method has relatively lower computational cost, while at the same maintain the benefits of simple transported PDF methods.
For high-pressure turbulent spray flames, modeling high-pressure turbulent spray, which can mimic the real flame in internal combustion engines, is a challenging task. Unlike the modeling for stationary turbulent flame, high-pressure turbulent spray flames require accurate high-temporal and spatial resolution. The task is even more difficult due to the transient ignition characteristics involving multiple combustion modes: premixed, mixing-controlled, and kinetically-controlled combustion. In direct-injection engines, the two-phase process of the spray increases the modeling complexity. Furthermore, limited reliable detailed data from experiments for real engine configurations limits the understanding of turbulent spray inside the cylinder.
It is important to remember that most of the diesel combustions are modeled based on measurements at atmospheric pressure [176,177,178,179]. In reality, the combustion process occurs at high-pressure involving spray breakup, droplet evaporation, two-phase heat and mass transfer, chemical kinetics, flame propagation, and other transport properties. This obviously requires a different approach as opposed to the near-atmospheric modeling method, considering the turbulent-chemistry interaction to realistically model the turbulent spray.
Therefore, in order to produce realistic predictions for combustion and emission characteristics under such conditions, the model should be supported by a number of experimental results. The experimentation should be conducted to represent detailed thermochemical process to fulfill the prerequisite of advanced combustion of diesel technology. The Engine Combustion Network (ECN) from Sandia National Laboratories, provides the reliable experimental database performed in practical diesel engines. This is carried out to bridge the gap in modeling between the laboratory-scale flame and real modern diesel engines.
Note that the engine performance is considerably responsive to slight variations in the physical state of combustible mixture. Therefore, accurate modeling of the injector and spray behavior are needed to achieve realistic CFD predictions. The spray near the injector, in particular, is of great concern when the injected liquid fuel undergoes a series of instable processes resulted from complex liquid-gas phase interactions. This region often referred to as primary atomization region, is the dense-spray area and has been studied extensively.
Still, comprehensive physical descriptions that affect the formation and growth of the spray have not yet been fully known. Furthermore, the lack of reliable experimental techniques that can sufficiently describe its physical characteristics has restricted the theoretical understanding of the spray formation. Several modeling findings, however, were performed to describe the spray behavior statistically. Yet, no experimental results exist to validate these models. To solve this problem, most modeling strategies have to tune the model coefficients to match measurements further downstream of the injector. In this region, often referred to dilute-spray region, the physical behavior of the spray is relatively easy to understand due to its regularity.
A number of Eulerian- as well as Lagrangian-based methods were successfully employed to model the spray behavior in dilute sprays [180,181,182,183,184]. Such methods are also able to model the two-phase interaction involving droplets (liquid) and surrounding in-cylinder gas. Of the two methods, Lagrangian-based dispersed-phase is more popular for diesel spray modeling. Most of the Lagrangian models analyze the droplet evaporation by assuming the droplet properties to be uniform and the droplet internal circulation to be considerably fast. The interactions between the dispersed-phase droplets and strong turbulent flow within the combustion chamber are important factors to control the evaporation as well as heat transfer rates to and from the droplets. Therefore, assuming the droplets to have uniform properties that change with time is a reasonable approach in the dilute-spray region.
Multi-component droplet evaporation has also been paid attention due to growing interest in the engine development that operate on multi fuels. Initially, multi-component vaporization studies estimated the droplet evaporation as a single spray component of vaporization sequence controlled by volatility differences. Later, it was found that for multi-fuels consisting of different mass diffusivities, the evaporation rates were influenced by each spray component and limited by the low-diffusivity component. In general, the droplet evaporates due to combined effects of both volatility and mass diffusion of each spray components. However, when the mass diffusivity drops to zero, the vaporization rates are no longer subject to volatility of the components.
Most diesel spray modeling neglects the diffusive transport effect inside the droplets. While this assumption is widely accepted in conventional diesel engine condition, in-depth consideration should be taken into account if this assumption to be applied in a low-pressure low-temperature diesel engine. In such condition, the evaporation rate is relatively low so that wall film formation could be formed. As a consequence, the transient heating of the wall film and differential mass diffusion of the liquid components can significantly change the evaporation rates, thus altering the combustion as well as emission behaviors.
All in all, although the advancement of the rapid computational method has significantly improved the numerical predictions of spray behavior, a thorough understanding of the fundamental physical mechanism remains the key factor to the research in engine combustion. The physical phenomena inside the cylinder should be able to be modeled with simple, yet realistic flame configurations without using complex geometries. After being compared with the experimental results, the CFD model could hopefully accurately represent real engine combustion.

6. Optical Measurement Systems

Diesel spray phenomena in the combustion process is difficult to measure. It is considerably challenging to capture spray behavior [185]. In order to gain a deeper comprehension of diesel spray and fuel-air mixing phenomena, several optical techniques have been introduced. The optical diagnostic aims to investigate the spray behavior using non-intrusive method [186]. This method is generally categorized into two types: photography and non-photography. The photography technique intends to observe the macroscopic characteristics such as fuel distribution, spray length, and angle, whereas the non-photographic technique seeks to study the microscopic characteristics.
Specifically, based on the working principle, optical diagnostic techniques are divided into two: traditional diagnostic techniques based on conventional optics and laser-based diagnostic techniques [187]. Figure 6 shows the laser absorption-scattering (LAS) technique basic principle. When the two laser beams go all the way through the spray, the attenuation effect decreases the intensities. In the ultraviolet image, the lower intensity is caused by scattering and absorption of the liquid as well as absorption of the vapor where the impact of the liquid absorption is insignificant [60,188,189].
The diagnostic techniques provide substantial information about the nature of the fuel flow or combustion under high temperature and pressure conditions, thus providing insights into the chemical and physical mechanisms of spray both on macro- and micro-scales. Various optical diagnostic methods have been adopted in previous studies, such as Mie scattering [191,192], Rayleigh scattering [193,194], Raman scattering [195,196], shadow graphy (SG) [191,197], phase doppler interferometry (PDI) [198], Schlieren photography [191,199], laser-induced (exciplex) fluorescence (LIF/LIEF) [123,200], and the combined method [201,202]. Table 1 shows the optical diagnostic techniques used for several phenomena capturing.
The distribution of liquid and vapor phases in diesel spray has gained great attention [126,189,203,204,205]. In the last two decades, a new technique called particle image velocimetry (PIV) has successfully advanced the research in mixture formation [206,207,208]. PIV has revealed that the nearby airflow can be classified into three different regions according to the flow property. In the first region (head vortex zone), the gas is initially pushed apart by the head front of the spray. The gas is subsequently recirculated in the recirculation zone (region 2). Finally, in the last region, region 3 (near quasi static zone), the gas is entrained into the following spray zone.

7. Conclusions

Diesel engines have relatively higher efficiency compared to the gasoline engine, achieved by a higher compression ratio by eliminating the throttling losses as commonly found in gasoline engines. This superior benefit of high efficiency in diesel engines comes at the expense of the high level of emissions. The two most significant pollutants of diesel engines are NOx and PM. The latter stands for particulate matter and is commonly known as soot. Subject to engine load conditions, diesel engines produce a substantial amount of NOx. Several factors that determine the NOx formation include thermal NO, fuel-bound nitrogen, prompt NO, and N2O-intermediate. It occurs in a lean-fuel region, while soot emission, on the other hand, occurs in a rich-fuel region.
Diesel engine performance and pollutant emissions are the outcome of the chemical and physical developments taking place in the cylinder when fuel jets are ignited. Important in-cylinder processes include: the spatial and temporal development of liquid-fuel sprays; fuel vaporization, the mixing of vaporized fuel with in-cylinder gases; the low-temperature autoignition chemistry of fuel components; the high-temperature chemical kinetics of combustion and pollutant formation; as well as late-cycle mixing and burnout processes.
Diesel spray is a crucial event affecting the diesel engine performance, combustion, and emission characteristics. It has received considerable attention not only from automobile manufacturers, but also from academicians across the world. There are several major aspects of diesel spray, which have been considered the dominant factors, including injector cavitation, break-up process, and spray development. Spray cavitation suggest the importance of turbulence phenomenon caused by vapor bubble disruption in an injector, whereas the break-up process indicates the existence of a problem caused by liquid surface instability during the spray development. To observe these events, a number of measurement methods have been proposed, encouraging a new scientific achievement known as diesel spray laser diagnostics.
Many studies have been conducted focusing on spray atomization. Several factors are known to be responsible for the spray atomization. They were: (i) the aerodynamic forces; (ii) internal turbulence; (iii) boundary mutation at the exit of the nozzle; and (iv) internal flow cavitation. With the help of the optically-accessible-measurement technique, none of the factors mentioned above are found to influence the spray atomization independently. In fact, it is believed that the combination of some dominant factors is actually responsible collectively.
It is difficult to determine whether the cavitation has a constructive or adverse effect in engine performance. Although, the cavitation enhances the spray atomization, it also reduces the effective flow area (cross sectional area) which, in fact, deteriorates the combustion under large amount condition. Most researchers are, therefore, trying to reach an agreement on the effects of cavitation on engine performance. Additionally, aerodynamic forces play an important role as the spray exits the nozzle hole [31,40,41]. After the spray passes break-up length, aerodynamic forces cause the smaller droplets formation. This phenomenon is known as the secondary break-up. It relies greatly upon the ratio of aerodynamic and surface tension (as gas phase Weber number). The process of break up with low Weber number occurs in the downstream region, while that with high number takes place in upstream region. It is also important to remember that in diesel engine, spray penetration and angle are of significance on the utility of ambient gas and mixture formation rate. Overlong spray penetration causes several negative consequences such as higher hydrocarbon and carbon monoxide emissions, lower fuel economy, as well as more lubricant consumption due to wall wetting. On the other hand, short spray penetration is known to deteriorate the utilization of chamber gas.
Observing spray behavior in the diesel engine is a challenging task, requiring techniques that involve measurement in high-speed high-pressure and high-temperature condition. There are, however, three parameters to measure the structure of the spray in diesel engine: spray shape, microscopic, and macroscopic characteristics. Besides spray shape and macroscopic parameters, microscopic is of great importance in diesel spray/combustion. Even though numerous studies have investigated these microscopic factors, no definitive theoretical concept has been made regarding the droplet size of diesel spray. These microscopic factors include: (i) mean diameter, (ii) droplets size distribution, and (iii) spatial distribution. For mean diameter, the Sauter mean diameter (SMD) is the most extensively utilized definition. The Sauter mean diameter could indicate the spray average evaporation qualities. For droplets size distribution, smaller droplet size leads to faster evaporation. Yet, smaller size droplets cannot maintain its momentum. As a result, after losing momentum, it fails to accelerate the mixing between fuel and air. For spatial distribution, this important factor affects the quality of diesel spray combustion. Although a number of optical measurements as well as numerical simulations have been conducted in the last two decades, the established evaluation method for spatial distribution of diesel spray has not yet been settled. A combination between the spray’s angle, penetration, volume and droplets’ spatial number concentration remain to be used to characterize the spatial spray distribution.

Author Contributions

Conceptualization, D.W.D., P.A.P. and O.D.S.; methodology, P.A.P., M.E.M.S. and S.G.H.; software, M.M.A. and H.P.; validation, D.W.D., P.A.P. and S.G.H.; formal analysis, D.W.D., M.I. and O.D.S.; investigation, M.E.M.S. and D.C.; resources, M.M.A., S.G.H. and A.A.Y.; data curation, M.I. and M.M.A.; writing—original draft preparation, D.W.D., M.I., M.E.M.S. and S.G.H.; writing—review and editing, D.C., A.A.Y., H.P. and I.V.; visualization, D.W.D. and H.P.; supervision, M.E.M.S., H.P. and I.V.; project administration, O.D.S., D.C. and A.A.Y.; funding acquisition, D.W.D. and I.V. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thank Sampoerna University for the financial support to this work.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tian, J.; Yu, L.; Xue, R.; Zhuang, S.; Shan, Y. Global low-carbon energy transition in the post-COVID-19 era. Appl. Energy 2022, 307, 118205. [Google Scholar] [CrossRef]
  2. Zhu, Y.; Zhou, W.; Xia, C.; Hou, Q. Application and Development of Selective Catalytic Reduction Technology for Marine Low-Speed Diesel Engine: Trade-Off among High Sulfur Fuel, High Thermal Efficiency, and Low Pollution Emission. Atmosphere 2022, 13, 731. [Google Scholar] [CrossRef]
  3. Mohan, S.; Dinesha, P.; Kumar, S. NOx reduction behaviour in copper zeolite catalysts for ammonia SCR systems: A review. Chem. Eng. J. 2020, 384, 123253. [Google Scholar] [CrossRef]
  4. Damma, D.; Ettireddy, P.R.; Reddy, B.M.; Smirniotis, P.G. A Review of Low Temperature NH3-SCR for Removal of NOx. Catalysts 2019, 9, 349. [Google Scholar] [CrossRef] [Green Version]
  5. Bytnerowicz, A.; Omasa, K.; Paoletti, E. Integrated effects of air pollution and climate change on forests: A northern hemisphere perspective. Environ. Pollut. 2007, 147, 438–445. [Google Scholar] [CrossRef]
  6. Veza, I.; Said, M.F.M.; Latiff, Z.A. Improved Performance, Combustion and Emissions of SI Engine Fuelled with Butanol: A Review. Int. J. Automot. Mech. Eng. 2020, 17, 7648–7666. [Google Scholar] [CrossRef] [Green Version]
  7. Rusli, M.Q.; Said, M.F.M.; Sulaiman, A.M.; Roslan, M.F.; Veza, I.; Perang, M.R.M.; Lau, H.L.N.; Wafti, N.S.A. Performance and Emission Measurement of a Single Cylinder Diesel Engine Fueled with Palm Oil Biodiesel Fuel Blends. IOP Conf. Series Mater. Sci. Eng. 2021, 1068, 012020. [Google Scholar] [CrossRef]
  8. Fan, B.; Pan, J.; Yang, W.; Chen, W.; Bani, S. The influence of injection strategy on mixture formation and combustion process in a direct injection natural gas rotary engine. Appl. Energy 2017, 187, 663–674. [Google Scholar] [CrossRef]
  9. Chen, W.; Pan, J.; Yang, W.; Liu, Y.; Fan, B.; Lu, Y.; Otchere, P. Stratified combustion characteristics analysis and assisted-ignition strategy optimization in a natural gas blended diesel Wankel engine. Fuel 2021, 292, 120192. [Google Scholar] [CrossRef]
  10. GWang, G.; Yu, W.; Li, X.; Yang, R. Influence of fuel injection and intake port on combustion characteristics of controllable intake swirl diesel engine. Fuel 2020, 262, 116548. [Google Scholar]
  11. Cao, D.N.; Hoang, A.T.; Luu, H.Q.; Bui, V.G.; Tran, T.T.H. Effects of injection pressure on the NOx and PM emission control of diesel engine: A review under the aspect of PCCI combustion condition. Energy Sources Part A Recover. Util. Environ. Eff. 2020, 23, 2908–2920. [Google Scholar] [CrossRef]
  12. Sener, R.; Yangaz, M.U.; Gul, M.Z. Effects of injection strategy and combustion chamber modification on a single-cylinder diesel engine. Fuel 2020, 266, 117122. [Google Scholar] [CrossRef]
  13. Chen, H.; Su, X.; He, J.; Zhang, P.; Xu, H.; Zhou, C. Investigation on combustion characteristics of cyclopentanol/diesel fuel blends in an optical engine. Renew. Energy 2020, 167, 811–829. [Google Scholar] [CrossRef]
  14. Zhang, M.; Hong, W.; Xie, F.; Su, Y.; Liu, H.; Zhou, S. Combustion, performance and particulate matter emissions analysis of operating parameters on a GDI engine by traditional experimental investigation and Taguchi method. Energy Convers. Manag. 2018, 164, 344–352. [Google Scholar] [CrossRef]
  15. La Rocca, A.; Ferrante, A.; Haffner-Staton, E.; Cairns, A.; Weilhard, A.; Sans, V.; Carlucci, A.P.; Laforgia, D. Investigating the impact of copper leaching on combustion characteristics and particulate emissions in HPCR diesel engines. Fuel 2020, 263, 116719. [Google Scholar] [CrossRef]
  16. Dhahad, H.A.; Fayad, M.A.; Chaichan, M.T.; Jaber, A.A.; Megaritis, T. Influence of fuel injection timing strategies on performance, combustion, emissions and particulate matter characteristics fueled with rapeseed methyl ester in modern diesel engine. Fuel 2021, 306, 121589. [Google Scholar] [CrossRef]
  17. Attia, A.M.; Kulchitskiy, A.; Nour, M.; El-Seesy, A.I.; Nada, S.A. The influence of castor biodiesel blending ratio on engine performance including the determined diesel particulate matters composition. Energy 2022, 239, 121951. [Google Scholar] [CrossRef]
  18. Song, J.; Lee, Z.; Song, J.; Park, S. Effects of injection strategy and coolant temperature on hydrocarbon and particulate emissions from a gasoline direct injection engine with high pressure injection up to 50 MPa. Energy 2018, 164, 512–522. [Google Scholar] [CrossRef]
  19. Wang, Z.; Dai, X.; Li, F.; Li, Y.; Lee, C.-F.; Wu, H.; Li, Z. Nozzle internal flow and spray primary breakup with the application of closely coupled split injection strategy. Fuel 2018, 228, 187–196. [Google Scholar] [CrossRef]
  20. Chen, Q.; Wang, C.; Shao, K.; Liu, Y.; Chen, X.; Qian, Y. Analyzing the combustion and emissions of a DI diesel engine powered by primary alcohol (methanol, ethanol, n-butanol)/diesel blend with aluminum nano-additives. Fuel 2022, 328, 125222. [Google Scholar] [CrossRef]
  21. Ooi, J.B.; Chow, M.R.; Chee, K.M.; Pun, C.H.; Tran, M.-V.; Leong, J.C.K.; Lim, S. Effects of ethanol on the evaporation and burning characteristics of palm-oil based biodiesel droplet. J. Energy Inst. 2021, 98, 35–43. [Google Scholar]
  22. Wu, F.; Wang, H.; Yu, H.; Zang, X.; Pan, X.; Hua, M.; Jiang, J. Experimental study on the lower explosion limit and mechanism of methanol pre-mixed spray under negative pressure. Fuel 2022, 321, 124049. [Google Scholar] [CrossRef]
  23. Wang, Y.; Qi, C.; Ning, Y.; Lv, X.; Yu, X.; Yan, X.; Yu, J. Experimental determination of the lower flammability limit and limiting oxygen concentration of propanal/air mixtures under elevated temperatures and pressures. Fuel 2022, 326, 124882. [Google Scholar] [CrossRef]
  24. Wang, H.; Luo, K.; Hawkes, E.R.; Chen, J.H.; Fan, J. Turbulence, evaporation and combustion interactions in n-heptane droplets under high pressure conditions using DNS. Combust. Flame 2021, 225, 417–427. [Google Scholar] [CrossRef]
  25. Rajasegar, R.; Niki, Y.; Li, Z.; García-Oliver, J.M.; Musculus, M.P. Influence of pilot-fuel mixing on the spatio-temporal progression of two-stage autoignition of diesel-sprays in low-reactivity ambient fuel-air mixture. Proc. Combust. Inst. 2021, 38, 5741–5750. [Google Scholar] [CrossRef]
  26. Dec, J.E. A conceptual model of DL diesel combustion based on laser-sheet imaging. SAE Trans. 1997, 106, 1319–1348. [Google Scholar]
  27. Demarco, R.; Jerez, A.; Liu, F.; Chen, L.; Fuentes, A. Modeling soot formation in laminar coflow ethylene inverse diffusion flames. Combust. Flame 2021, 232, 111513. [Google Scholar] [CrossRef]
  28. Wang, L.-Y.; Chatterjee, S.; An, Q.; Steinberg, A.M.; Gülder, L. Soot formation and flame structure in swirl-stabilized turbulent non-premixed methane combustion. Combust. Flame 2019, 209, 303–312. [Google Scholar] [CrossRef]
  29. Li, B.; Zhang, D.; Liu, J.; Tian, Y.; Gao, Q.; Li, Z. A Review of Femtosecond Laser-Induced Emission Techniques for Combustion and Flow Field Diagnostics. Appl. Sci. 2019, 9, 1906. [Google Scholar] [CrossRef] [Green Version]
  30. Ruan, C.; Chen, F.; Cai, W.; Qian, Y.; Yu, L.; Lu, X. Principles of non-intrusive diagnostic techniques and their applications for fundamental studies of combustion instabilities in gas turbine combustors: A brief review. Aerosp. Sci. Technol. 2019, 84, 585–603. [Google Scholar] [CrossRef]
  31. Cheng, Q.; Ahmad, Z.; Kaario, O.; Vuorinen, V.; Larmi, M. Experimental study on tri-fuel combustion using premixed methane-hydrogen mixtures ignited by a diesel pilot. Int. J. Hydrog. Energy 2021, 46, 21182–21197. [Google Scholar] [CrossRef]
  32. Beatrice, C.; Denbratt, I.; Di Blasio, G.; Di Luca, G.; Ianniello, R.; Saccullo, M. Experimental Assessment on Exploiting Low Carbon Ethanol Fuel in a Light-Duty Dual-Fuel Compression Ignition Engine. Appl. Sci. 2020, 10, 7182. [Google Scholar] [CrossRef]
  33. Abdelaal, M.; Hegab, A. Combustion and emission characteristics of a natural gas-fueled diesel engine with EGR. Energy Convers. Manag. 2012, 64, 301–312. [Google Scholar] [CrossRef]
  34. Gehmlich, R.; Mueller, C.; Ruth, D.; Nilsen, C.; Skeen, S.; Manin, J. Using ducted fuel injection to attenuate or prevent soot formation in mixing-controlled combustion strategies for engine applications. Appl. Energy 2018, 226, 1169–1186. [Google Scholar] [CrossRef]
  35. Rao, L.; Zhang, Y.; Kim, D.; Su, H.C.; Kook, S.; Kim, K.S.; Kweon, C.-B. Effect of after injections on late cycle soot oxidation in a small-bore diesel engine. Combust. Flame 2018, 191, 513–526. [Google Scholar] [CrossRef]
  36. Fayad, M.A.; Al-Salihi, H.A.; Dhahad, H.A.; Mohammed, F.M.; Al-Ogidi, B.R. Effect of post-injection and alternative fuels on combustion, emissions and soot nanoparticles characteristics in a common-rail direct injection diesel engine. Energy Sources Part A Recovery Util. Environ. Eff. 2021, 1–15. [Google Scholar] [CrossRef]
  37. Gad, H.; Ibrahim, I.; Abdel-Baky, M.; El-Samed, A.A.; Farag, T. Experimental study of diesel fuel atomization performance of air blast atomizer. Exp. Therm. Fluid Sci. 2018, 99, 211–218. [Google Scholar] [CrossRef]
  38. Zhang, P.; Su, X.; Yi, C.; Chen, H.; Xu, H.; Geng, L. Spray, atomization and combustion characteristics of oxygenated fuels in a constant volume bomb: A review. J. Traffic Transp. Eng. Engl. Ed. 2020, 7, 282–297. [Google Scholar] [CrossRef]
  39. Xia, J.; Huang, Z.; Xu, L.; Ju, D.; Lu, X. Experimental study on spray and atomization characteristics under subcritical, transcritical and supercritical conditions of marine diesel engine. Energy Convers. Manag. 2019, 195, 958–971. [Google Scholar] [CrossRef]
  40. YSun, Y.; Guan, Z.; Hooman, K. Cavitation in Diesel Fuel Injector Nozzles and its Influence on Atomization and Spray. Chem. Eng. Technol. 2019, 42, 6–29. [Google Scholar]
  41. Anez, J.; Ahmed, A.; Hecht, N.; Duret, B.; Reveillon, J.; Demoulin, F. Eulerian–Lagrangian spray atomization model coupled with interface capturing method for diesel injectors. Int. J. Multiph. Flow 2019, 113, 325–342. [Google Scholar] [CrossRef] [Green Version]
  42. Kim, Y.-I.; Kim, S.; Yang, H.-M.; Lee, K.-Y.; Choi, Y.-S. Analysis of internal flow and cavitation characteristics for a mixed-flow pump with various blade thickness effects. J. Mech. Sci. Technol. 2019, 33, 3333–3344. [Google Scholar] [CrossRef]
  43. Cheng, H.Y.; Bai, X.R.; Long, X.P.; Ji, B.; Peng, X.X.; Farhat, M. Large eddy simulation of the tip-leakage cavitating flow with an insight on how cavitation influences vorticity and turbulence. Appl. Math. Model. 2020, 77, 788–809. [Google Scholar] [CrossRef]
  44. Mamaikin, D.; Knorsch, T.; Rogler, P.; Wensing, M. Experimental investigation of flow field and string cavitation inside a transparent real-size GDI nozzle. Exp. Fluids 2020, 61, 154. [Google Scholar] [CrossRef]
  45. Zhou, J.; Andersson, M. An analysis of surface breakup induced by laser-generated cavitation bubbles in a turbulent liquid jet. Exp. Fluids 2020, 61, 242. [Google Scholar] [CrossRef]
  46. Liu, Z.; Li, Z.; Liu, J.; Wu, J.; Yu, Y.; Ding, J. Numerical Study on Primary Breakup of Disturbed Liquid Jet Sprays Using a VOF Model and LES Method. Processes 2022, 10, 1148. [Google Scholar] [CrossRef]
  47. Trummler, T.; Schmidt, S.J.; Adams, N.A. Investigation of condensation shocks and re-entrant jet dynamics in a cavitating nozzle flow by Large-Eddy Simulation. Int. J. Multiph. Flow 2020, 125, 103215. [Google Scholar] [CrossRef] [Green Version]
  48. Chryssakis, C.A.; Assanis, D.N.; Tanner, F.X. Atomization Models. In Handbook of Atomization and Sprays: Theory and Applications; Ashgriz, N., Ed.; Springer: Boston, MA, USA, 2011; pp. 215–231. [Google Scholar]
  49. Alozie, N.S.; Ganippa, L.C. Diesel Exhaust Emissions and Mitigations; IntechOpen: London, UK, 2019. [Google Scholar]
  50. Zhou, L.; Zhao, W.; Luo, K.H.; Wei, H.; Xie, M. Spray–turbulence–chemistry interactions under engine-like conditions. Prog. Energy Combust. Sci. 2021, 86, 100939. [Google Scholar] [CrossRef]
  51. Chintagunti, S.J.; Kalwar, A.; Kumar, D.; Agarwal, A.K. Spray Chamber Designs and Optical Techniques for Fundamental Spray Investigations. In Novel Internal Combustion Engine Technologies for Performance Improvement and Emission Reduction; Springer: Singapore, 2021; pp. 105–144. [Google Scholar]
  52. Hamdi, F.; Agrebi, S.; Idrissi, M.S.; Mondo, K.; Labiadh, Z.; Sadiki, A.; Chrigui, M. Impact of Spray Cone Angle on the Performances of Methane/Diesel RCCI Engine Combustion under Low Load Operating Conditions. Entropy 2022, 24, 650. [Google Scholar] [CrossRef]
  53. Zhan, C.; Luo, H.; Chang, F.; Nishida, K.; Ogata, Y.; Tang, C.; Feng, Z.; Huang, Z. Experimental study on the droplet characteristics in the spray tip region: Comparison between the free and impinging spray. Exp. Therm. Fluid Sci. 2020, 121, 110288. [Google Scholar] [CrossRef]
  54. Peraza, J.E.; Salvador, F.J.; Gimeno, J.; Ruiz, S. ECN Spray D visualization of the spray interaction with a transparent wall under engine-like conditions. Part I: Non-reactive impinging spray. Fuel 2022, 307, 121699. [Google Scholar] [CrossRef]
  55. Peraza, J.E.; Payri, R.; Gimeno, J.; Martí-Aldaraví, P. ECN Spray D visualization of the spray interaction with a transparent wall under engine-like conditions, Part II: Impinging spray combustion. Fuel 2022, 308, 121964. [Google Scholar] [CrossRef]
  56. Aizawa, T.; Kinoshita, T.; Akiyama, S.; Shinohara, K.; Miyagawa, Y. Infrared high-speed thermography of combustion chamber wall impinged by diesel spray flame. Int. J. Engine Res. 2021, 23, 1116–1130. [Google Scholar] [CrossRef]
  57. Yang, K.; Nishida, K.; Yamakawa, H. Effect of split injection ratio on combustion process of diesel spray into two-dimensional piston cavity. Fuel 2020, 260, 116316. [Google Scholar] [CrossRef]
  58. Mahmud, R.; Kurisu, T.; Nishida, K.; Ogata, Y.; Kanzaki, J.; Akgol, O. Effects of injection pressure and impingement distance on flat-wall impinging spray flame and its heat flux under diesel engine-like condition. Adv. Mech. Eng. 2019, 11, 1687814019862910. [Google Scholar] [CrossRef] [Green Version]
  59. Bothell, J.K.; Machicoane, N.; Li, D.; Morgan, T.B.; Aliseda, A.; Kastengren, A.L.; Heindel, T.J. Comparison of X-ray and optical measurements in the near-field of an optically dense coaxial air-assisted atomizer. Int. J. Multiph. Flow 2020, 125, 103219. [Google Scholar] [CrossRef] [Green Version]
  60. Chen, R.; Nishida, K.; Shi, B. Quantitative investigation on the spray mixture formation for ethanol-gasoline blends via UV–Vis dual-wavelength laser absorption scattering (LAS) technique. Fuel 2019, 242, 425–437. [Google Scholar] [CrossRef]
  61. Li, Y.; Wang, Z.; Kong, Q.; Li, B.; Wang, H. Sulfur dioxide absorption by charged droplets in electrohydrodynamic atomization. Int. Commun. Heat Mass Transf. 2022, 137, 106275. [Google Scholar] [CrossRef]
  62. Jia, H.; Wei, Z.; Yin, B.; Liu, Z. Analysis of elliptical diesel nozzle spray dynamics using a one-way coupled spray model. Int. J. Engine Res. 2021, 14680874211063352. [Google Scholar] [CrossRef]
  63. Xie, K.; Cui, Y.; Qiu, X.; Wang, J. Experimental study on flame characteristics and air entrainment of diesel horizontal spray burners at two different atmospheric pressures. Energy 2020, 211, 118906. [Google Scholar] [CrossRef]
  64. Wei, Y.; Li, T.; Zhou, X.; Zhang, Z. Time-resolved measurement of the near-nozzle air entrainment of high-pressure diesel spray by high-speed micro-PTV technique. Fuel 2020, 268, 117343. [Google Scholar] [CrossRef]
  65. Wei, Y.; Li, T.; Chen, R.; Zhou, X.; Zhang, Z.; Wang, X. Measurement and modeling of the near-nozzle ambient gas entrainment of high-pressure diesel sprays. Fuel 2022, 310, 122373. [Google Scholar] [CrossRef]
  66. Santos, E.G.; Shi, J.; Gavaises, M.; Soteriou, C.; Winterbourn, M.; Bauer, W. Investigation of cavitation and air entrainment during pilot injection in real-size multi-hole diesel nozzles. Fuel 2020, 263, 116746. [Google Scholar] [CrossRef]
  67. Hiroyasu, H.; Arai, M. Structures of fuel sprays in diesel engines. SAE Trans. 1990, 99, 1050–1061. [Google Scholar]
  68. Machicoane, N.; Bothell, J.K.; Li, D.; Morgan, T.B.; Heindel, T.J.; Kastengren, A.L.; Aliseda, A. Synchrotron radiography characterization of the liquid core dynamics in a canonical two-fluid coaxial atomizer. Int. J. Multiph. Flow 2019, 115, 1–8. [Google Scholar] [CrossRef]
  69. Kong, Q.; Yang, S.; Wang, Q.; Wang, Z.; Dong, Q.; Wang, J. Dynamics of electrified jets in electrohydrodynamic atomization. Case Stud. Therm. Eng. 2022, 29, 101725. [Google Scholar] [CrossRef]
  70. Berni, F.; Sparacino, S.; Riccardi, M.; Cavicchi, A.; Postrioti, L.; Borghi, M.; Fontanesi, S. A zonal secondary break-up model for 3D-CFD simulations of GDI sprays. Fuel 2022, 309, 122064. [Google Scholar] [CrossRef]
  71. Koukouvinis, P.; Vidal-Roncero, A.; Rodriguez, C.; Gavaises, M.; Pickett, L. High pressure/high temperature multiphase simulations of dodecane injection to nitrogen: Application on ECN Spray-A. Fuel 2020, 275, 117871. [Google Scholar] [CrossRef]
  72. Pielecha, I. The influence of petrol injection parameters on the structure of geometry of fuel spray injected from outward-opening injectors. Fuel 2018, 222, 64–73. [Google Scholar] [CrossRef]
  73. Payri, R.; Gimeno, J.; Martí-Aldaraví, P.; Martínez, M. Transient nozzle flow analysis and near field characterization of gasoline direct fuel injector using Large Eddy Simulation. Int. J. Multiph. Flow 2022, 148, 103920. [Google Scholar] [CrossRef]
  74. Cui, J.; Lai, H.; Feng, K.; Ma, Y. Quantitative analysis of the minor deviations in nozzle internal geometry effect on the cavitating flow. Exp. Therm. Fluid Sci. 2018, 94, 89–98. [Google Scholar] [CrossRef]
  75. Monieta, J.; Kasyk, L. Optimization of Design and Technology of Injector Nozzles in Terms of Minimizing Energy Losses on Friction in Compression Ignition Engines. Appl. Sci. 2021, 11, 7341. [Google Scholar] [CrossRef]
  76. Piscaglia, F.; Giussani, F.; Hèlie, J.; Lamarque, N.; Aithal, S. Vortex Flow and Cavitation in Liquid Injection: A Comparison between High-Fidelity CFD Simulations and Experimental Visualizations on Transparent Nozzle Replicas. Int. J. Multiph. Flow 2021, 138, 103605. [Google Scholar] [CrossRef]
  77. Yang, S.; Ma, Z.; Li, X.; Hung, D.L.; Xu, M. A review on the experimental non-intrusive investigation of fuel injector phase changing flow. Fuel 2020, 259, 116188. [Google Scholar] [CrossRef]
  78. Wang, C.; Adams, M.; Jin, T.; Sun, Y.; Röll, A.; Luo, F.; Gavaises, M. An analytical model of diesel injector’s needle valve eccentric motion. Int. J. Engine Res. 2022, 23, 469–481. [Google Scholar] [CrossRef]
  79. Gavaises, M.; Murali-Girija, M.; Rodriguez, C.; Koukouvinis, P.; Gold, M.; Pearson, R. Numerical simulation of fuel dribbling and nozzle wall wetting. Int. J. Engine Res. 2022, 23, 132–149. [Google Scholar] [CrossRef]
  80. Torres-Garcia, M.; García-Martín, J.F.; Aguilar, F.J.J.-E.; Barbin, D.F.; Alvarez-Mateos, P. Vegetable oils as renewable fuels for power plants based on low and medium speed diesel engines. J. Energy Inst. 2020, 93, 953–961. [Google Scholar] [CrossRef]
  81. Lee, Z.; Kim, T.; Park, S.; Park, S. Review on spray, combustion, and emission characteristics of recent developed direct-injection spark ignition (DISI) engine system with multi-hole type injector. Fuel 2020, 259, 116209. [Google Scholar] [CrossRef]
  82. Li, H.; Rutland, C.J.; Pérez, F.E.H.; Im, H.G. Large-eddy spray simulation under direct-injection spark-ignition engine-like conditions with an integrated atomization/breakup model. Int. J. Engine Res. 2021, 22, 731–754. [Google Scholar] [CrossRef] [Green Version]
  83. Guo, G.; He, Z.; Wang, Q.; Lai, M.-C.; Zhong, W.; Guan, W.; Wang, J. Numerical investigation of transient hole-to-hole variation in cavitation regimes inside a multi-hole diesel nozzle. Fuel 2021, 287, 119457. [Google Scholar] [CrossRef]
  84. Chouak, M.; Dufresne, L.; Seers, P. Large eddy simulation of a double-injection cycle and the impact of the needle motion on the sac-volume flow characteristics of a single-orifice diesel injector. Int. J. Engine Res. 2021, 22, 2464–2476. [Google Scholar] [CrossRef]
  85. Julien, L. Characteristics of Diesel Sprays at High Temperatures and Pressures. Ph.D. Thesis, The University of Brighton, Brighton, UK, 2006. [Google Scholar]
  86. Kale, R.; Banerjee, R. Experimental investigation on GDI spray behavior of isooctane and alcohols at elevated pressure and temperature conditions. Fuel 2019, 236, 1–12. [Google Scholar] [CrossRef]
  87. Zigan, L.; Schmitz, I.; Flügel, A.; Wensing, M.; Leipertz, A. Structure of evaporating single- and multicomponent fuel sprays for 2nd generation gasoline direct injection. Fuel 2011, 90, 348–363. [Google Scholar] [CrossRef]
  88. Han, J.-S.; Lu, P.-H.; Xie, X.-B.; Lai, M.-C.; Henein, N.A. Investigation of diesel spray primary break-up and development for different nozzle geometries. SAE Trans. 2002, 111, 2528–2548. [Google Scholar]
  89. Yu, S.; Yin, B.; Bi, Q.; Jia, H.; Chen, C. Effects of gasoline and ethanol on inner flows and swallowtail-like spray behaviors of elliptical GDI injector. Fuel 2021, 294, 120543. [Google Scholar] [CrossRef]
  90. Som, S.; Longman, D.; Ramírez, A.; Aggarwal, S. A comparison of injector flow and spray characteristics of biodiesel with petrodiesel. Fuel 2010, 89, 4014–4024. [Google Scholar] [CrossRef]
  91. Battistoni, M.; Grimaldi, C.N. Numerical analysis of injector flow and spray characteristics from diesel injectors using fossil and biodiesel fuels. Appl. Energy 2012, 97, 656–666. [Google Scholar] [CrossRef]
  92. Kostas, J.; Honnery, D.; Soria, J. A correlation image velocimetry-based study of high-pressure fuel spray tip evolution. Exp. Fluids 2011, 51, 667–678. [Google Scholar] [CrossRef]
  93. Wu, G.; Zhou, X.; Li, T. Temporal Evolution of Split-Injected Fuel Spray at Elevated Chamber Pressures. Energies 2019, 12, 4284. [Google Scholar] [CrossRef] [Green Version]
  94. Klein-Douwel, R.J.H.; Frijters, P.J.M.; Somers, L.M.T.; de Boer, W.A.; Baert, R.S.G. Macroscopic diesel fuel spray shadowgraphy using high speed digital imaging in a high pressure cell. Fuel 2007, 86, 1994–2007. [Google Scholar] [CrossRef]
  95. Chen, L.; Li, G.; Huang, D.; Zhang, Z.; Lu, Y.; Yu, X.; Roskilly, A.P. Experimental and numerical study on the initial tip structure evolution of diesel fuel spray under various injection and ambient pressures. Energy 2019, 186, 115867. [Google Scholar] [CrossRef]
  96. Ghasemi, A.; Li, X.; Hong, Z.; Yun, S. Breakup mechanisms in air-assisted atomization of highly viscous pyrolysis oils. Energy Convers. Manag. 2020, 220, 113122. [Google Scholar] [CrossRef]
  97. Zhang, W.; Liu, H.; Liu, C.; Jia, M.; Xi, X. Numerical investigation into primary breakup of diesel spray with residual bubbles in the nozzle. Fuel 2019, 250, 265–276. [Google Scholar] [CrossRef]
  98. Sykes, D.; Turner, J.; Stetsyuk, V.; de Sercey, G.; Gold, M.; Pearson, R.; Crua, C. Quantitative characterisations of spray deposited liquid films and post-injection discharge on diesel injectors. Fuel 2021, 289, 119833. [Google Scholar] [CrossRef]
  99. Fauchais, P.L.; Heberlein, J.V.; Boulos, M.I. Overview of thermal spray. In Thermal Spray Fundamentals; Springer: Berlin/Heidelberg, Germany, 2014; pp. 17–72. [Google Scholar]
  100. Zhang, X.; Ranjith, P.G. Experimental investigation of effects of CO2 injection on enhanced methane recovery in coal seam reservoirs. J. CO2 Util. 2019, 33, 394–404. [Google Scholar] [CrossRef]
  101. Yip, H.L.; Srna, A.; Yuen, A.C.Y.; Kook, S.; Taylor, R.A.; Yeoh, G.H.; Medwell, P.R.; Chan, Q.N. A Review of Hydrogen Direct Injection for Internal Combustion Engines: Towards Carbon-Free Combustion. Appl. Sci. 2019, 9, 4842. [Google Scholar] [CrossRef]
  102. Boretti, A. Advances in Diesel-LNG Internal Combustion Engines. Appl. Sci. 2020, 10, 1296. [Google Scholar] [CrossRef] [Green Version]
  103. McTaggart-Cowan, G.; Mann, K.; Huang, J.; Singh, A.; Patychuk, B.; Zheng, Z.X.; Munshi, S. Direct Injection of Natural Gas at up to 600 Bar in a Pilot-Ignited Heavy-Duty Engine. SAE Int. J. Engines 2015, 8, 981–996. [Google Scholar] [CrossRef]
  104. Hamzehloo, A.; Aleiferis, P. Gas dynamics and flow characteristics of highly turbulent under-expanded hydrogen and methane jets under various nozzle pressure ratios and ambient pressures. Int. J. Hydrog. Energy 2016, 41, 6544–6566. [Google Scholar] [CrossRef] [Green Version]
  105. Yu, S.; Yin, B.; Deng, W.; Jia, H.; Ye, Z.; Xu, B.; Xu, H. Internal flow and spray characteristics for elliptical orifice with large aspect ratio under typical diesel engine operation conditions. Fuel 2018, 228, 62–73. [Google Scholar] [CrossRef]
  106. Zhao, J.; Grekhov, L.; Yue, P. Limit of Fuel Injection Rate in the Common Rail System under Ultra-High Pressures. Int. J. Automot. Technol. 2020, 21, 649–656. [Google Scholar] [CrossRef]
  107. Wu, H.; Zhang, F.; Zhang, Z.; Gao, H. Experimental investigation on the spray characteristics of a self-pressurized hollow cone injector. Fuel 2020, 272, 117710. [Google Scholar] [CrossRef]
  108. Yu, S.; Yin, B.; Deng, W.; Jia, H.; Ye, Z.; Xu, B.; Xu, H. Experimental study on the diesel and biodiesel spray characteristics emerging from equilateral triangular orifice under real diesel engine operation conditions. Fuel 2018, 224, 357–365. [Google Scholar] [CrossRef]
  109. Yu, S.; Yin, B.; Deng, W.; Jia, H.; Ye, Z.; Xu, B.; Xu, H. Experimental study on the spray characteristics discharging from elliptical diesel nozzle at typical diesel engine conditions. Fuel 2018, 221, 28–34. [Google Scholar] [CrossRef]
  110. Dhanji, M.; Zhao, H. Investigations of split injection properties on the spray characteristics using a solenoid high-pressure injector. Int. J. Engine Res. 2022, 23, 262–284. [Google Scholar] [CrossRef]
  111. Aleiferis, P.; Papadopoulos, N. Heat and mass transfer effects in the nozzle of a fuel injector from the start of needle lift to after the end of injection in the presence of fuel dribble and air entrainment. Int. J. Heat Mass Transf. 2021, 165, 120576. [Google Scholar] [CrossRef]
  112. Brulatout, J.; Garnier, F.; Seers, P. Interaction between a diesel-fuel spray and entrained air with single- and double-injection strategies using large eddy simulations. Propuls. Power Res. 2020, 9, 37–50. [Google Scholar] [CrossRef]
  113. Zhan, C.; Feng, Z.; Zhang, M.; Tang, C.; Huang, Z. Experimental investigation on effect of ethanol and di-ethyl ether addition on the spray characteristics of diesel/biodiesel blends under high injection pressure. Fuel 2018, 218, 1–11. [Google Scholar] [CrossRef]
  114. Algayyim, S.J.M.; Wandel, A.P. Macroscopic and microscopic characteristics of biofuel spray (biodiesel and alcohols) in CI engines: A review. Fuel 2021, 292, 120303. [Google Scholar] [CrossRef]
  115. Luo, H.; Nishida, K.; Uchitomi, S.; Ogata, Y.; Zhang, W.; Fujikawa, T. Microscopic behavior of spray droplets under flat-wall impinging condition. Fuel 2018, 219, 467–476. [Google Scholar] [CrossRef]
  116. Hawi, M.; Kosaka, H.; Sato, S.; Nagasawa, T.; Elwardany, A.; Ahmed, M. Effect of injection pressure and ambient density on spray characteristics of diesel and biodiesel surrogate fuels. Fuel 2019, 254, 115674. [Google Scholar] [CrossRef]
  117. Zhang, P.; Su, X.; Chen, H.; Geng, L.; Zhao, X. Assessing fuel properties effects of 2,5-dimethylfuran on microscopic and macroscopic characteristics of oxygenated fuel/diesel blends spray. Sci. Rep. 2020, 10, 1427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Liu, J.; Feng, L.; Wang, H.; Zheng, Z.; Chen, B.; Zhang, D.; Yao, M. Spray characteristics of gasoline/PODE and diesel/PODE blends in a constant volume chamber. Appl. Therm. Eng. 2019, 159, 113850. [Google Scholar] [CrossRef]
  119. Ulu, A.; Yildiz, G.; Özkol, Ü.; Rodriguez, A.D. Experimental investigation of spray characteristics of ethyl esters in a constant volume chamber. Biomass Convers. Biorefin. 2022, 23, 1–18. [Google Scholar] [CrossRef] [PubMed]
  120. Shi, C.; Bu, S.; Zhang, L.; Yuan, H.; Xu, W.; Liu, L.; Zhang, Z. Experimental and numerical investigation on the evaporation performance of a cyclone-type spray desalination chamber. Desalination 2019, 467, 125–135. [Google Scholar] [CrossRef]
  121. Sun, Y.; Alkhedhair, A.M.; Guan, Z.; Hooman, K. Numerical and experimental study on the spray characteristics of full-cone pressure swirl atomizers. Energy 2018, 160, 678–692. [Google Scholar] [CrossRef]
  122. Badawy, T.; Xu, H.; Li, Y. Macroscopic spray characteristics of iso-octane, ethanol, gasoline and methanol from a multi-hole injector under flash boiling conditions. Fuel 2022, 307, 121820. [Google Scholar] [CrossRef]
  123. Feng, L.; Sun, X.; Pan, X.; Yi, W.; Cui, Y.; Wang, Y.; Wen, M.; Ming, Z.; Liu, H.; Yao, M. Gasoline spray characteristics using a high pressure common rail diesel injection system by the method of laser induced exciplex fluorescence. Fuel 2021, 302, 121174. [Google Scholar] [CrossRef]
  124. Shi, Z.; Lee, C.-F.; Wu, H.; Wu, Y.; Zhang, L.; Liu, F. Optical diagnostics of low-temperature ignition and combustion characteristics of diesel/kerosene blends under cold-start conditions. Appl. Energy 2019, 251, 113307. [Google Scholar] [CrossRef]
  125. Cao, T.; He, Z.; Si, Z.; El-Seesy, A.I.; Guan, W.; Zhou, H.; Wang, Q. Optical experimental study on cavitation development with different patterns in diesel injector nozzles at different fuel temperatures. Exp. Fluids 2020, 61, 185. [Google Scholar] [CrossRef]
  126. Shi, Z.; Lee, C.-F.; Wu, H.; Li, H.; Wu, Y.; Zhang, L.; Bo, Y.; Liu, F. Effect of injection pressure on the impinging spray and ignition characteristics of the heavy-duty diesel engine under low-temperature conditions. Appl. Energy 2020, 262, 114552. [Google Scholar] [CrossRef]
  127. Yan, J.; Gao, S.; Liu, W.; Chen, T.; Lee, T.H.; Lee, C.-F. Experimental study of flash boiling spray with isooctane, hexane, ethanol and their binary mixtures. Fuel 2021, 292, 120415. [Google Scholar] [CrossRef]
  128. Han, M.; Gordon, R.L.; Talei, M.; Lacey, J.S. Ignition of dense, inhomogeneous fuel sprays at elevated pressures and temperatures. Fuel 2022, 321, 123853. [Google Scholar] [CrossRef]
  129. Wu, S.; Yang, S.; Wooldridge, M.; Xu, M. Experimental study of the spray collapse process of multi-hole gasoline fuel injection at flash boiling conditions. Fuel 2019, 242, 109–123. [Google Scholar] [CrossRef]
  130. Zhang, Z.; Lu, Y.; Qian, Z.; Roskilly, A.P. Spray and engine performance of cerium oxide nanopowder and carbon nanotubes modified alternative fuel. Fuel 2022, 320, 123952. [Google Scholar] [CrossRef]
  131. Reuss, D.L.; Kim, N.; Sjöberg, M. The influence of intake flow and coolant temperature on gasoline spray morphology during early-injection DISI engine operation. Int. J. Engine Res. 2022, 14680874221104301. [Google Scholar] [CrossRef]
  132. Duy, T.-N.; Nguyen, V.-T.; Phan, T.-H.; Hwang, H.-S.; Park, W.-G. Numerical analysis of ventilated cavitating flow around an axisymmetric object with different discharged temperature conditions. Int. J. Heat Mass Transf. 2022, 197, 123338. [Google Scholar] [CrossRef]
  133. Podbevšek, D.; Lokar, Ž.; Podobnikar, J.; Petkovšek, R.; Dular, M. Experimental evaluation of methodologies for single transient cavitation bubble generation in liquids. Exp. Fluids 2021, 62, 167. [Google Scholar] [CrossRef]
  134. Sanli, H.; Alptekin, E.; Canakci, M. Using low viscosity micro-emulsification fuels composed of waste frying oil-diesel fuel-higher bio-alcohols in a turbocharged-CRDI diesel engine. Fuel 2022, 308, 121966. [Google Scholar] [CrossRef]
  135. Bari, S.; Hossain, S.; Saad, I. A review on improving airflow characteristics inside the combustion chamber of CI engines to improve the performance with higher viscous biofuels. Fuel 2019, 264, 116769. [Google Scholar] [CrossRef]
  136. Hamid, M.F.; Idroas, M.Y.; Sa’ad, S.; Yew Heng, T.; Che Mat, S.; Zainal Alauddin, Z.A.; Shamsuddin, K.A.; Shuib, R.K.; Abdullah, M.K. Numerical investigation of fluid flow and in-cylinder air flow characteristics for higher viscosity fuel applications. Processes 2020, 8, 439. [Google Scholar] [CrossRef] [Green Version]
  137. Cui, Y.; Liu, H.; Geng, C.; Tang, Q.; Feng, L.; Wang, Y.; Yi, W.; Zheng, Z.; Yao, M. Optical diagnostics on the effects of fuel properties and coolant temperatures on combustion characteristic and flame development progress from HCCI to CDC via PPC. Fuel 2020, 269, 117441. [Google Scholar] [CrossRef]
  138. Venu, H.; Raju, V.D.; Lingesan, S.; Soudagar, M.E.M. Influence of Al2O3nano additives in ternary fuel (diesel-biodiesel-ethanol) blends operated in a single cylinder diesel engine: Performance, Combustion and Emission Characteristics. Energy 2021, 215, 119091. [Google Scholar] [CrossRef]
  139. Jhalani, A.; Sharma, D.; Soni, S.L.; Sharma, P.K.; Sharma, S. A comprehensive review on water-emulsified diesel fuel: Chemistry, engine performance and exhaust emissions. Environ. Sci. Pollut. Res. 2019, 26, 4570–4587. [Google Scholar] [CrossRef]
  140. Khandavalli, S.; Sharma-Nene, N.; Kabir, S.; Sur, S.; Rothstein, J.P.; Neyerlin, K.C.; Mauger, S.A.; Ulsh, M. Toward Optimizing Electrospun Nanofiber Fuel Cell Catalyst Layers: Polymer–Particle Interactions and Spinnability. ACS Appl. Polym. Mater. 2021, 3, 2374–2384. [Google Scholar] [CrossRef]
  141. Geo, V.E.; Prabhu, C.; Thiyagarajan, S.; Maiyalagan, T.; Aloui, F. Comparative analysis of various techniques to improve the performance of novel wheat germ oil—An experimental study. Int. J. Hydrog. Energy 2020, 45, 5745–5756. [Google Scholar]
  142. Broumand, M.; Albert-Green, S.; Yun, S.; Hong, Z.; Thomson, M.J. Spray combustion of fast pyrolysis bio-oils: Applications, challenges, and potential solutions. Prog. Energy Combust. Sci. 2020, 79, 100834. [Google Scholar] [CrossRef]
  143. Dafsari, R.A.; Lee, H.J.; Han, J.; Park, D.-C.; Lee, J. Viscosity effect on the pressure swirl atomization of an alternative aviation fuel. Fuel 2019, 240, 179–191. [Google Scholar] [CrossRef]
  144. Thongchai, S.; Lim, O. Influence of Biodiesel Blended in Gasoline-Based Fuels on Macroscopic Spray Structure from a Diesel Injector. Int. J. Automot. Technol. 2019, 20, 701–711. [Google Scholar] [CrossRef]
  145. Liu, F.; Li, Z.; Wang, Z.; Dai, X.; He, X.; Lee, C.-F. Microscopic study on diesel spray under cavitating conditions by injecting fuel into water. Appl. Energy 2018, 230, 1172–1181. [Google Scholar] [CrossRef]
  146. Das, S.K.; Kim, K.; Lim, O. Experimental study on non-vaporizing spray characteristics of biodiesel-blended gasoline fuel in a constant volume chamber. Fuel Process. Technol. 2018, 178, 322–335. [Google Scholar] [CrossRef]
  147. Ashikhmin, A.E.; Khomutov, N.A.; Piskunov, M.V.; Yanovsky, V.A. Secondary Atomization of a Biodiesel Micro-Emulsion Fuel Droplet Colliding with a Heated Wall. Appl. Sci. 2020, 10, 685. [Google Scholar] [CrossRef] [Green Version]
  148. Patiño-Camino, R.; Cova-Bonillo, A.; Lapuerta, M.; Rodríguez-Fernández, J.; Segade, L. Surface tension of diesel-alcohol blends: Selection among fundamental and empirical models. Fluid Phase Equilibria 2022, 555, 113363. [Google Scholar] [CrossRef]
  149. Singh, G.; Pham, P.; Kourmatzis, A.; Masri, A. Effect of electric charge and temperature on the near-field atomization of diesel and biodiesel. Fuel 2019, 241, 941–953. [Google Scholar] [CrossRef]
  150. Panchasara, H.; Ashwath, N. Effects of Pyrolysis Bio-Oils on Fuel Atomisation—A Review. Energies 2021, 14, 794. [Google Scholar] [CrossRef]
  151. Pham, P.X.; Nguyen, K.T.; Pham, T.V.; Nguyen, V.H. Biodiesels Manufactured from Different Feedstock: From Fuel Properties to Fuel Atomization and Evaporation. ACS Omega 2020, 5, 20842–20853. [Google Scholar] [CrossRef]
  152. Antonov, D.V.; Kuznetsov, G.V.; Strizhak, P.A.; Fedorenko, R.M. Micro-explosion of droplets containing liquids with different viscosity, interfacial and surface tension. Chem. Eng. Res. Des. 2020, 158, 129–147. [Google Scholar] [CrossRef]
  153. Park, S.; Park, K. Principles and droplet size distributions of various spraying methods: A review. J. Mech. Sci. Technol. 2022, 36, 4033–4041. [Google Scholar] [CrossRef]
  154. Emerson, P.; Crockett, J.; Maynes, D. Thermal atomization during droplet impingement on superhydrophobic surfaces: Influence of Weber number and micropost array configuration. Int. J. Heat Mass Transf. 2021, 164, 120559. [Google Scholar] [CrossRef]
  155. Shlegel, N.; Tkachenko, P.; Strizhak, P. Influence of viscosity, surface and interfacial tensions on the liquid droplet collisions. Chem. Eng. Sci. 2020, 220, 115639. [Google Scholar] [CrossRef]
  156. Hoang, A.T.; Le, A.T.; Pham, V.V. A core correlation of spray characteristics, deposit formation, and combustion of a high-speed diesel engine fueled with Jatropha oil and diesel fuel. Fuel 2019, 244, 159–175. [Google Scholar] [CrossRef]
  157. Zhang, G.; Shi, P.; Luo, H.; Ogata, Y.; Nishida, K. Investigation on fuel adhesion characteristics of wall-impingement spray under cross-flow conditions. Fuel 2022, 320, 123925. [Google Scholar] [CrossRef]
  158. Zhuang, Y.; Chi, H.; Huang, Y.; Teng, Q.; He, B.; Chen, W.; Qian, Y. Investigation of water spray evolution process of port water injection and its effect on engine performance. Fuel 2020, 282, 118839. [Google Scholar] [CrossRef]
  159. Sathiyamoorthi, R.; Sankaranarayanan, G.; Munuswamy, D.B.; Devarajan, Y. Experimental study of spray analysis for Palmarosa biodiesel-diesel blends in a constant volume chamber. Environ. Prog. Sustain. Energy 2021, 40, e13696. [Google Scholar] [CrossRef]
  160. Biswal, A.; Kale, R.; Balusamy, S.; Banerjee, R.; Kolhe, P. Lemon peel oil as an alternative fuel for GDI engines: A spray characterization perspective. Renew. Energy 2019, 142, 249–263. [Google Scholar] [CrossRef]
  161. Chen, Y.; Liu, S.; Guo, X.; Jia, C.; Huang, X.; Wang, Y.; Huang, H. Experimental Research on the Macroscopic and Microscopic Spray Characteristics of Diesel-PODE3-4 Blends. Energies 2021, 14, 5559. [Google Scholar] [CrossRef]
  162. Wu, H.; Zhang, F.; Zhang, Z. Fundamental spray characteristics of air-assisted injection system using aviation kerosene. Fuel 2021, 286, 119420. [Google Scholar] [CrossRef]
  163. Suraj, C.; Sudarshan, G.; Anand, K.; Sundararajan, T. Effects of autooxidation on the fuel spray characteristics of Karanja biodiesel. Biomass Bioenergy 2021, 149, 106084. [Google Scholar] [CrossRef]
  164. Yan, J.; Gao, S.; Zhao, W.; Lee, T.H.; Lee, C.-F. Experimental study of sprays with isooctane, hexane, ethanol and their binary mixtures under different flash boiling intensities. Int. J. Heat Mass Transf. 2021, 179, 121715. [Google Scholar] [CrossRef]
  165. Mei, S.S.; Rahman, A.A.A.; Abidin, M.S.Z.; Mazlan, N.M. d2 Law and Penetration Length of Jatropha and Camelina Bio-Synthetic Paraffinic Kerosene Spray Characteristics at Take-Off, Top of Climb and Cruise. Aerospace 2021, 8, 249. [Google Scholar]
  166. Wang, Y.; Zhuang, Y.; Yao, M.; Qin, Y.; Zheng, Z. An experimental investigation into the soot particle emissions at early injection timings in a single-cylinder research diesel engine. Fuel 2022, 316, 123288. [Google Scholar] [CrossRef]
  167. Hwang, J.; Weiss, L.; Karathanassis, I.K.; Koukouvinis, P.; Pickett, L.M.; Skeen, S.A. Spatio-temporal identification of plume dynamics by 3D computed tomography using engine combustion network spray G injector and various fuels. Fuel 2020, 280, 118359. [Google Scholar] [CrossRef]
  168. Yi, P.; Li, T.; Wei, Y.; Zhou, X. Experimental and numerical investigation of low sulfur heavy fuel oil spray characteristics under high temperature and pressure conditions. Fuel 2021, 286, 119327. [Google Scholar] [CrossRef]
  169. Zhou, Z.-F.; Liang, L.; Murad, S.H.M.; Camm, J.; Davy, M. Investigation of fuel volatility on the heat transfer dynamics on piston surface due to the pulsed spray impingement. Int. J. Heat Mass Transf. 2021, 170, 121008. [Google Scholar] [CrossRef]
  170. He, X.; Li, Y.; Liu, C.; Sjöberg, M.; Vuilleumier, D.; Liu, F.; Yang, Q. Characteristics of spray and wall wetting under flash-boiling and non-flashing conditions at varying ambient pressures. Fuel 2020, 264, 116683. [Google Scholar] [CrossRef]
  171. Bao, J.; Qu, P.; Wang, H.; Zhou, C.; Zhang, L.; Shi, C. Implementation of various bowl designs in an HPDI natural gas engine focused on performance and pollutant emissions. Chemosphere 2022, 303, 135275. [Google Scholar] [CrossRef] [PubMed]
  172. Hoang, A.T. Combustion behavior, performance and emission characteristics of diesel engine fuelled with biodiesel containing cerium oxide nanoparticles: A review. Fuel Process. Technol. 2021, 218, 106840. [Google Scholar] [CrossRef]
  173. Rajak, U.; Nashine, P.; Verma, T.N.; Pugazhendhi, A. Performance, combustion and emission analysis of microalgae Spirulina in a common rail direct injection diesel engine. Fuel 2019, 255, 115855. [Google Scholar] [CrossRef]
  174. Hoang, A.T.; Le, A.T. A review on deposit formation in the injector of diesel engines running on biodiesel. Energy Sources Part A Recover. Util. Environ. Eff. 2019, 41, 584–599. [Google Scholar] [CrossRef]
  175. Jenny, P.; Roekaerts, D.; Beishuizen, N. Modeling of turbulent dilute spray combustion. Prog. Energy Combust. Sci. 2012, 38, 846–887. [Google Scholar] [CrossRef]
  176. Huang, H.; Lv, D.; Zhu, J.; Zhu, Z.; Chen, Y.; Pan, Y.; Pan, M. Development of a new reduced diesel/natural gas mechanism for dual-fuel engine combustion and emission prediction. Fuel 2019, 236, 30–42. [Google Scholar] [CrossRef]
  177. Pang, K.M.; Jangi, M.; Bai, X.-S.; Schramm, J.; Walther, J.H.; Glarborg, P. Effects of ambient pressure on ignition and flame characteristics in diesel spray combustion. Fuel 2019, 237, 676–685. [Google Scholar] [CrossRef]
  178. Finesso, R.; Hardy, G.; Mancarella, A.; Marello, O.; Mittica, A.; Spessa, E. Real-time simulation of torque and nitrogen oxide emissions in an 11.0 L heavy-duty diesel engine for model-based combustion control. Energies 2019, 12, 460. [Google Scholar] [CrossRef] [Green Version]
  179. Rubio, J.A.P.; Vera-García, F.; Grau, J.H.; Cámara, J.M.; Hernandez, D.A. Marine diesel engine failure simulator based on thermodynamic model. Appl. Therm. Eng. 2018, 144, 982–995. [Google Scholar] [CrossRef]
  180. Yu, F. Numerical Studies of Nuclear Containment Spray Process by Stochastic Field Method and CGCFD Approach. Ph.D. Thesis, Institut für Thermische Energietechnik und Sicherheit, Karlsruhe, Germany, 2020. [Google Scholar]
  181. Boel, E.; Koekoekx, R.; Dedroog, S.; Babkin, I.; Vetrano, M.R.; Clasen, C.; Van den Mooter, G. Unraveling Particle Formation: From Single Droplet Drying to Spray Drying and Electrospraying. Pharmaceutics 2020, 12, 625. [Google Scholar] [CrossRef]
  182. Sharma, M.; Goyal, D.K.; Kaushal, G.; Grover, N.K.; Bansal, A.; Goyal, K. CFD and experimental study of slurry erosion wear in Hydro-machinery. Mater. Today Proc. 2022, 62, 7581–7594. [Google Scholar] [CrossRef]
  183. Han, S.; Zhang, R.; Song, Y.; Xing, J.; Zhou, L.; Li, L.; Zhang, H.; Du, X. Numerical study of swirl cooling enhancement by adding mist to air: Effects of droplet diameter and mist concentration. Appl. Therm. Eng. 2022, 211, 118475. [Google Scholar] [CrossRef]
  184. Venkatachalam, P.; Sahu, S.; Anupindi, K. Investigation of cross-stream spray injection and wall impingement in a circular channel for SCR application. Therm. Sci. Eng. Prog. 2022, 32, 101229. [Google Scholar] [CrossRef]
  185. Fansler, T.D.; Parrish, S. Spray measurement technology: A review. Meas. Sci. Technol. 2014, 26, 012002. [Google Scholar] [CrossRef]
  186. Leipertz, A.; Wensing, M. Modern optical diagnostics in engine research. J. Phys. Conf. Ser. 2007, 85, 012001. [Google Scholar] [CrossRef]
  187. Xu, H.E.; Yue, W.U.; Xiao, M.A.; Yanfei, L.I.; Yunliang, Q.I.; Zechang, L.I.U.; Yifan, X.U.; Yang, Z.H.O.U.; Xiongwei, L.I.; Cong, L.I.U.; et al. A review of optical diagnostic platforms and techniques applied in internal combustion engines. Shiyan Liuti Lixue (J. Exp. Fluid Mech.) 2020, 34, 1–52. [Google Scholar]
  188. Qi, W.; Zhang, Y. Quantitative measurement of binary-component fuel vapor distributions via laser absorption and scattering imaging. Appl. Phys. B 2019, 125, 127. [Google Scholar] [CrossRef]
  189. Zhou, Y.; Wei, Z.; Zhu, Q.; Cao, Y.; Zhang, Y. Quantitative characterization on cyclic variation of mixture formation for flash boiling sprays. Energy 2022, 257, 124808. [Google Scholar] [CrossRef]
  190. Jin, Y.; Wu, Q.; Zhai, C.; Kim, J.; Luo, H.-L.; Ogata, Y.; Nishida, K. Evaporating characteristics of diesel sprays under split-injection condition with a negative dwell time. Energetic Mater. Front. 2021, 2, 265–271. [Google Scholar] [CrossRef]
  191. Kim, D.; Park, S.S.; Bae, C. Schlieren, Shadowgraph, Mie-scattering visualization of diesel and gasoline sprays in high pressure/high temperature chamber under GDCI engine low load condition. Int. J. Automot. Technol. 2018, 19, 1–8. [Google Scholar] [CrossRef]
  192. Mounaïm-Rousselle, C.; Pajot, O. Droplet sizing by Mie scattering interferometry in a spark ignition engine. Part. Part. Syst. Charact. Meas. Descr. Part. Prop. Behav. Powders Other Disperse Syst. 1999, 16, 160–168. [Google Scholar] [CrossRef]
  193. Idicheria, C.A.; Pickett, L.M. Quantitative mixing measurements in a vaporizing diesel spray by Rayleigh imaging. SAE Trans. 2007, 116, 490–504. [Google Scholar]
  194. Pickett, L.M.; Genzale, C.L.; Manin, J.; Malbec, L.-M.; Hermant, L. Measurement uncertainty of liquid penetration in evaporating diesel sprays. In Proceedings of the ILASS Americas, 23rd Annual Conference on Liquid Atomization and Spray Systems, Ventura, CA, USA, 15–18 May 2011. [Google Scholar]
  195. Egermann, J.; Taschek, M.; Leipertz, A. Spray/wall interaction influences on the diesel engine mixture formation process investigated by spontaneous Raman scattering. Proc. Combust. Inst. 2002, 29, 617–623. [Google Scholar] [CrossRef]
  196. Egermann, J.; Göttler, A.; Leipertz, A. Application of spontaneous Raman scattering for studying the diesel mixture formation process under near-wall conditions. SAE Trans. 2001, 110, 2182–2188. [Google Scholar]
  197. Emberson, D.; Ihracska, B.; Imran, S.; Diez, A. Optical characterization of Diesel and water emulsion fuel injection sprays using shadowgraphy. Fuel 2016, 172, 253–262. [Google Scholar] [CrossRef] [Green Version]
  198. Gupta, J.G.; Agarwal, A.K. Macroscopic and Microscopic Spray Characteristics of Diesel and Karanja Biodiesel Blends; SAE Technical Paper0148-7191; SAE International: Warrendale, PA, USA, 2016. [Google Scholar]
  199. Payri, R.; Salvador, F.; Bracho, G.; Viera, A. Differences between single and double-pass schlieren imaging on diesel vapor spray characteristics. Appl. Therm. Eng. 2017, 125, 220–231. [Google Scholar] [CrossRef]
  200. Bruneaux, G. Mixing process in high pressure diesel jets by normalized laser induced exciplex fluorescence: Part i: Free jet. SAE Trans. 2005, 114, 1444–1461. [Google Scholar]
  201. Adam, A.; Leick, P.; Bittlinger, G.; Schulz, C. Visualization of the evaporation of a diesel spray using combined Mie and Rayleigh scattering techniques. Exp. Fluids 2009, 47, 439–449. [Google Scholar] [CrossRef] [Green Version]
  202. Liu, R.; Huang, L.; Feng, M.; Ju, D.; Ma, Z.; Lu, X. Schlieren and Mie Scattering Visualization of Liquid and Vapor Phase Behavior for Large Nozzle Diameter Injectors Under Marine Diesel Engine Conditions. SSRN 2022. [Google Scholar] [CrossRef]
  203. Markov, V.; Sa, B.; Devyanin, S.; Grekhov, L.; Neverov, V.; Zhao, J. Numerical analysis of injection and spray characteristics of diesel fuel and rapeseed oil in a diesel engine. Case Stud. Therm. Eng. 2022, 35, 102129. [Google Scholar] [CrossRef]
  204. Guan, W.; He, Z.; Zhang, L.; El-Seesy, A.I.; Wen, L.; Zhang, Q.; Yang, L. Effect of asymmetric structural characteristics of multi-hole marine diesel injectors on internal cavitation patterns and flow characteristics: A numerical study. Fuel 2021, 283, 119324. [Google Scholar] [CrossRef]
  205. Zhao, Z.; Zhu, X.; Naber, J.; Lee, S.-Y. Assessment of impinged flame structure in high-pressure direct diesel injection. Int. J. Engine Res. 2020, 21, 391–405. [Google Scholar] [CrossRef]
  206. Jardón-Pérez, L.E.; González-Rivera, C.; Trápaga-Martínez, G.; Amaro-Villeda, A.; Ramírez-Argáez, M.A. Experimental Study of Mass Transfer Mechanisms for Solute Mixing in a Gas-Stirred Ladle Using the Particle Image Velocimetry and Planar Laser-Induced Fluorescence Techniques. Steel Res. Int. 2021, 92, 2100241. [Google Scholar] [CrossRef]
  207. Bilsky, A.V.; Gobyzov, O.A.; Markovich, D.M. Evolution and recent trends of particle image velocimetry for an aerodynamic experiment (review). Thermophys. Aeromech. 2020, 27, 1–22. [Google Scholar] [CrossRef]
  208. Jena, A.; Singh, A.P.; Agarwal, A.K. Challenges and Opportunities of Particle Imaging Velocimetry as a Tool for Internal Combustion Engine Diagnostics. In Novel Internal Combustion Engine Technologies for Performance Improvement and Emission Reduction; Springer: Singapore, 2021; pp. 43–77. [Google Scholar]
Figure 1. Diesel engine four strokes process.
Figure 1. Diesel engine four strokes process.
Sustainability 14 15902 g001
Figure 2. Combustion model of direct injection diesel, reproduced from [26].
Figure 2. Combustion model of direct injection diesel, reproduced from [26].
Sustainability 14 15902 g002
Figure 3. Mechanism of droplet break-up, reproduced from [48].
Figure 3. Mechanism of droplet break-up, reproduced from [48].
Sustainability 14 15902 g003
Figure 4. Illustration of diesel spray macroscopically, reproduced from [67].
Figure 4. Illustration of diesel spray macroscopically, reproduced from [67].
Sustainability 14 15902 g004
Figure 5. Three types of injector nozzle in diesel engines: (a) mini-sac volume; (b) micro-sac volume; and (c) valve covered orifice (VCO).
Figure 5. Three types of injector nozzle in diesel engines: (a) mini-sac volume; (b) micro-sac volume; and (c) valve covered orifice (VCO).
Sustainability 14 15902 g005
Figure 6. LAS technique basic principle, reproduced from [190].
Figure 6. LAS technique basic principle, reproduced from [190].
Sustainability 14 15902 g006
Table 1. Optical diagnostics techniques for fuel spray phenomena.
Table 1. Optical diagnostics techniques for fuel spray phenomena.
TechniqueApplicationPhenomena CapturingReferences
Mie ScatteringDiesel, gasoline engineLiquid distribution of sprayKim et al. [191]
Mie ScatteringGasoline engineDroplet sizeMounaïm-Rousselle
and Pajot [192]
Raman ScatteringDiesel engineSpray-wall interactionEgermann et al. [195]
Raman ScatteringDiesel engineSpray-wall interaction (near wall condition)Egermann et al. [196]
Raylaigh ScatteringDiesel engineVaporizing diesel sprayIdicheria and Pickett [193]
Raylaigh ScatteringDiesel engineLiquid penetration Pickett et al. [194]
Schlieren Diesel, gasoline engineVapor distributionKim et al. [191]
SchlierenDiesel engineVaporizing diesel sprayPayri et al. [199]
ShadowgraphyDiesel, gasoline engineVapor distribution (a mare shadow)Kim et al. [191]
ShadowgraphyDiesel engineDiesel—water emulsion fuel injection spraysEmberson et al. [197]
Phase doppler interferometry (PDI)Diesel engineMacroscopic and microscopic sprayGupta and Agrawal [198]
Laser-induced (exciplex) fluorescence (LIF/LIEF)Gasoline engineMixing processBruneaux [200]
Combined (Mie-Rayleigh)Diesel engineEvaporation of diesel sprayAdam et al. [201]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Djamari, D.W.; Idris, M.; Paristiawan, P.A.; Abbas, M.M.; Samuel, O.D.; Soudagar, M.E.M.; Herawan, S.G.; Chandran, D.; Yusuf, A.A.; Panchal, H.; et al. Diesel Spray: Development of Spray in Diesel Engine. Sustainability 2022, 14, 15902. https://doi.org/10.3390/su142315902

AMA Style

Djamari DW, Idris M, Paristiawan PA, Abbas MM, Samuel OD, Soudagar MEM, Herawan SG, Chandran D, Yusuf AA, Panchal H, et al. Diesel Spray: Development of Spray in Diesel Engine. Sustainability. 2022; 14(23):15902. https://doi.org/10.3390/su142315902

Chicago/Turabian Style

Djamari, Djati Wibowo, Muhammad Idris, Permana Andi Paristiawan, Muhammad Mujtaba Abbas, Olusegun David Samuel, Manzoore Elahi M. Soudagar, Safarudin Gazali Herawan, Davannendran Chandran, Abdulfatah Abdu Yusuf, Hitesh Panchal, and et al. 2022. "Diesel Spray: Development of Spray in Diesel Engine" Sustainability 14, no. 23: 15902. https://doi.org/10.3390/su142315902

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