0% found this document useful (0 votes)
74 views

Injection Rate Model

This document discusses an experimental and modeling study of liquid fuel injection and combustion in diesel engines with a common rail injection system. The study investigated the influence of injection pressure and timing on the injection rate profile. A new injection model was developed that can simulate the instantaneous fuel injection rate and duration for different injection pressures and durations. The model was shown to accurately replicate experimental injection rate profiles and predict engine performance under various operating conditions.

Uploaded by

aravind
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
74 views

Injection Rate Model

This document discusses an experimental and modeling study of liquid fuel injection and combustion in diesel engines with a common rail injection system. The study investigated the influence of injection pressure and timing on the injection rate profile. A new injection model was developed that can simulate the instantaneous fuel injection rate and duration for different injection pressures and durations. The model was shown to accurately replicate experimental injection rate profiles and predict engine performance under various operating conditions.

Uploaded by

aravind
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 66

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/327394671

Experimental and modeling study of liquid fuel injection and combustion in


diesel engines with a common rail injection system

Article  in  Applied Energy · November 2018


DOI: 10.1016/j.apenergy.2018.08.104

CITATIONS READS

63 861

6 authors, including:

xu Leilei Xue-Song Bai


Lund University Lund University
21 PUBLICATIONS   304 CITATIONS    250 PUBLICATIONS   4,859 CITATIONS   

SEE PROFILE SEE PROFILE

Ming Jia Yong Qian


Dalian University of Technology Shanghai Jiao Tong University
240 PUBLICATIONS   3,823 CITATIONS    105 PUBLICATIONS   1,577 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Characterization of metal particle combustion using optical diagnostics View project

Investigation of the uncertainty quantification of skeletal mechanism View project

All content following this page was uploaded by xu Leilei on 22 October 2019.

The user has requested enhancement of the downloaded file.


Experimental and modeling study of liquid fuel injection and combustion
in diesel engines with a common rail injection system
Leilei Xu a,b, Xue-Song Bai b, Ming Jia c, Yong Qian a, Xinqi Qiao a, Xingcai Lu a, *

a Key Lab. for Power Machinery and Engineering of M. O. E., Shanghai Jiao Tong University, 200240, P. R. China

b Division of Fluid Mechanics, Lund University, 221 00 Lund, Sweden

c School of Energy and Power Engineering, Dalian University of Technology, PR China

Corresponding author: E-mail address: [email protected]. Tel.: +86-21-34206039; Fax: +86-21-34205949.

Note:You can download the injection rate model from the link:

https://www.researchgate.net/publication/333949095_injection_rate_modelzip

Abstract

Fuel injection is one of the most important processes in compression-ignition internal combustion engines owing to

its significant impact on the exhaust emissions and thermal efficiency. In this study, experiments were carried out to

investigate the influence of injection pressure and injection timing on the temporal evolution of the injection rate and

injection duration in a specially designed experiment rig equipped with a common rail injection system. It is well

known that the injection signal from the electronic control unit (ECU) of the injection system, which is often the only

injection information available in engine operation and experiments, gives little information about the actual injection

rate profile. It is shown in the present experiments that the actual injection duration is usually longer than the

energizing time (ET). The time delay between the actual injection of the fuel and the ECU signal is about 0.3 – 0.4

ms, and the time delay appears to be insensitive to the injector geometry and injection pressure condition. The

injection process can be characterized as five stages, a fast injector valve opening stage, a slow valve opening stage,

a valve fully open stage, followed by a slow valve closing stage, and finally a rapid valve closing stage. It is found
that the first stage, the fast valve opening stage, is insensitive to the injection pressure and injector nozzle diameter;

however, the peak injection rate is a strong function of these parameters. The second and the third stage may not

appear with a short injection duration. A new injection model was developed for the common rail injection system,

which was capable of simulating the instantaneous fuel injection rate and injection duration for a range of injection

pressure and injection duration. The model was shown to be able to replicate the experimental injection rate profile

of the present experiments and experiments found in the literature for common rail injection system. The new

injection model was applied to predict the effect of injection pressure and injection duration on the performance of a

diesel engine under various engine speed and load conditions. The new injection model was shown to be able to

describe the injection mass flow rate, which eventually leads to a reasonably good prediction of the variations of the

spray development, in-cylinder pressure, heat release rate, and emissions.

Keywords: Common Rail Injection System; New injection model; Injection rate; CFD; Diesel engine

1. Introduction

In order to simultaneously enhance the thermal efficiency of internal combustion engines and to meet more

stringent emission regulations, novel combustion strategies have been developed in the past decades, e.g.,

homogeneous charge compression ignition (HCCI) [1], premixed charge compression ignition (PCCI) [2], partially

premixed combustion (PPC) [3, 4], reactivity controlled compression ignition (RCCI) [5, 6], modulated kinetics (MK)

[7], and low-temperature combustion (LTC) [8]. These novel incylinder strategies are aimed to minimize the use of

expensive after-treatment technologies [9, 10] that can reduce exhaust emissions from internal combustion engines

but with significant extra cost and complexity to an engine system. In most of these novel conceptual engines the

fuel is injected directly into the cylinder and the combustion process is greatly affected by the liquid fuel spray [11].

A desirable liquid fuel injection is to break the primary liquid fuel into smaller droplets that allows for fast evaporation

1
rate. The fuel spray quality governs the fuel/air mixing and heat transfer process in the combustion chamber, which

in turn dictates the quality of the combustion process. The spray process is typically initiated as a result of high-

pressure liquid fuel injection from several nozzles. The liquid inside the injector undergoes physical processes such

as cavitation [12], gas-phase bubble formation [13], and turbulent flows [14]. The complex multiple phase fluid can

possibly implode as it travels internally in the injector and when it is ejected into the chamber [15]. The injection rate

profile, which describes the velocity/mass flow rate of the liquid fuel as a function of time during the injection period,

is one of the most important parameters of the injector, which affects the evaporation rate, the fuel/air mixing, and

hence the combustion process and pollutant emissions. Therefore, injection system and nozzle geometry were

extensively studied in engine research community and in engine industry to achieve a desirable fuel injection rate.

Kashdan et al. [16] investigated the effect of injection rate on the combustion process using a single injection

strategy. It was shown that the injection rate has a noticeable influence on the main heat release; a slower injection

rate leads to a higher peak heat release rate and a higher level of engine noise. Tay et al. [17] numerically investigated

the effect of ramp and triangular injection rate profiles on the spray and combustion process of a diesel engine.

Although the start of injection and the duration of injection were kept constant, the injection rate profile was shown

to have an obvious effect on the ignition delay time. Furthermore, the injection rate profile could be used to control

the combustion phasing and combustion duration. Boggavarapu et al. [18] compared the impact of different fuel

injection rate profiles on a medium-duty diesel engine. It was reported that a slow increase in the injection rate in the

early injection process led to a reduction in nitrogen oxides (NOx) and indicated mean effective pressure (IMEP),

but the injection rate profile showed almost no effect on soot emissions. Using boot injection rate profiles, Mohan et

al. [19] varied the boot length (long, medium and short boot length) and boot pressure (low, medium and high boot

pressure), and it was found that a trade-off between NOx and soot emissions were obtained for the injection rate

profile with varying boot length and boot pressure. Rottmann et al. [20] investigated different injection rate profiles,

2
including rectangular (Common-Rail type), ramp and boot profiles, at high exhaust gas recirculation (EGR) rates.

For a constant start of injection and a constant EGR rate, the boot injection rate profile was shown to reduce NOx

emission and reduce combustion noise without the penalties in other emissions and engine efficiency. The

experimental study by Tanabe et al. [21] showed that the combustion process could be well controlled by the injection

rate profiles, and the tradeoff of NOx-fuel consumption and NOx-PM (particulate matter) emissions could be

substantially improved by optimizing the injection rate profile. Wakisaka et al. [22] estimated the influence of fuel

injection rate profile and injection pressure on diesel combustion using an electronically controlled fuel injection

system. The study revealed that the increased injection rate affected the temporal evolution of the spray, and higher

injection pressure enhanced spray development, particularly for the spray width. Macian et al. [23] observed that the

injection rate profiles had a substantial impact on the ignition event, including the initial ignition location in the

cylinder and the subsequent premixed combustion phase. Agarwal et al. [24] numerically investigated the effects of

the in-nozzle orifice parameters and injection parameters on the evolution of fuel spray in a constant-volume spray

chamber. The results showed that the nozzle-hole design was a significant parameter for obtaining the optimum spray

behavior in direct injection (DI) combustion engines.

It should be pointed out that despite the importance of the injection rate profile in the operation of DI engines,

measurement of the injection rate profile is seldom done. In practical engine operation, the fuel injection is controlled

by an electric control unit (ECU). The electric current signal profile is a signature of the injection rate profile; however,

it is not the same as the injection rate profile. In fact, there is a time delay in the start of the injection and the start of

the electric signal. It is important to develop a suitable model that can predict the injected fuel mass and the actual

injection rate profile from the electric signal, not only for the quantification of the impact of injection strategies on

the combustion and emission process in DI engines, but also for the predictive computational fluid dynamics (CFD)

modeling of DI engines. In most CFD studies of DI engines, effort has been made on the development and validation

3
of spray models [25-28], while less effort has been put on the accurate description of the inject rate profiles. It has

been found that a successful CFD run using codes, e.g., CONVERGE [29], OpenFoam [30] and KIVA [31], etc.,

requires an accurate description of the initial and boundary conditions, and detailed information of the injection rate

profiles.

Several injection rate models have been proposed [31-34]. In the KIVA code, three pulsed injection rate models

are currently available [31]. The first is the continuous spray model that may be useful in situations that fuel is

continuously injected into the combusting environment. The second is the half-sine wave or square wave model that

may be useful for theoretical studies. The third is the tabular input model that is based on the actual injection rate

profile from the experiments. The injection rate profile was defined over the injection duration specified by an

injection duration parameter. A detailed injection rate model based on the underlying physics of the process should

contain three sub-systems [32-34]: a high pressure pump sub-model, a common rail sub-model and an injector sub-

model. In these sub-models, proper treatment of the stiffness of the injector needle assembly and the discharge coefficients

of the different flow restrictions in the injector are very difficult but imperative for an accurate description of injection

process. Due to the complexity of the nozzle flow and the multi-body dynamics of the injection system, a detailed injection

rate model is not available.

The aim of this paper is to propose a robust injection rate model for the quantification of the impact of injection

strategies on DI engine combustion and emissions, and for the accurate description of the injection mass flow rate

for CFD simulation of DI engines. The model is based on the injection information typically available in the engine

experiments. Although the fuel injection rate profile is hardly available in the engine experiment, the total injected

fuel mass per cycle can be measured easily in the experiment. The new injection rate model is based on the injected

mass per cycle, the rail pressure, and the electric signal profile, taking into account the time delay between the start

of injection and the start of the electric signal. To validate the model, injection rate profiles in specially designed

4
experimental rigs in our laboratory and in the literature are simulated. The new injection model was shown to be

able to replicate the experimental results. The injection model was then evaluated further in real engine applications,

in which the effect of the injection pressure and duration on the performance of a diesel engine was studied.

2. Experiment

2.1. Experimental setup

Figure 1 shows the experimental setup for the measurements of the injection rate profiles in a test bench. The rig

is made up of three parts: a common rail injection system, an electronic control unit, ECU, and a data acquisition

system. Table 1 shows the specification of the main test apparatus in the experiment. ECU was used to control the

rail pressure, the injection pulse profiles, and the fuel supply. A mono-injection qualifier (EFS8246) supplied by the

French company EFS was used to measure the instantaneous injection rate. The electric current controlled by the

ECU, which is often referred to as the injector energizing current, was measured using a current sensor. A fuel pressure

sensor installed at the injector inlet was used to monitor the fuel pressure trace. The data acquisition device was used

to store the fuel injection rate, the energizing current and the inlet pressure of the injector. More detailed specifications

of the experiment can be found in Ref. [35].

The mono-injection qualifier was shown in Fig. 2. The flow meter measures the stroke-by-stroke instantaneous

injection volume flow rate. Depending on the mechanical configuration used, the maximum measured instantaneous

volume flow ranged from 100 mm3 to 600 mm3 with an accuracy of 0.1% FS (Full Scale). The oil cooling circuit

ensures the cooling of the measurement sensor and enables a greater accuracy. The used oil was an electric bench

adjustment oil with the thermostat set at 40 oC.

5
In the experiment, the rate of injection (ROI) of a 7-hole injector with 155 degree included spray angle (hereafter

referred to as Injector A) was measured for the energizing time from 0.5 ms to 1.8 ms and the injection pressure from

80 to 140 MPa. Here, the time span of energizing current signal is referred to as the energizing time, ET. The total

fuel mass flow rate was measured for at least 2000 injection cycles in order to control the random error due to the

cycle-to-cycle variation of the injection rate.

In addition to the current experiments, the experiment of Lee [36] is also considered for model validation. The

injector in the experiment of Lee has 8 nozzle holes and 120 degree included spray angle (the case is referred to as

Injector B). The ROIs were measured from 0.2 ms to 1.4 ms energizing time and the injection pressure from 50 to

150 MPa. Table 2 summarizes the nozzle parameters for the two injector cases.

2.2. Experimental Results

Figure 3a shows the energizing current signal and the rate of injection, ROI, of the case with 120 MPa injection

pressure and 0.9 ms energizing time for different injection cycles. In order to control the measurement uncertainty,

an average of 20 injections was calculated and shown in the figure. The scaling factor, which converts a signal profile

to the mass flow rate of injection, was determined according to the ratio of the total injection fuel mass to the area of

the signal-time plot. Figure 3b shows the average mass flow rate with error-bars representing the maximum and

minimum values in the 20 injections measured in the experiment. The deviations in the initial and end stages are

about 15% due to the violent fluctuation of pressure in the injector noted during the transient lift and fall of the needle.

The average deviation during the whole injection duration is about 4.47%, which is within an acceptable range.

As shown in Fig. 4a, a time delay can be seen between the start of current signal and the actual start of injection

(SOI) of the fuel due to the elasticity of the control plunger/needle. The injection pressure oscillates with time. The

6
pressure decrease during the energizing time (ET) is originated from the pressure wave in the control volume above

the control plunger when the ball valve lifts. When the injection event ends, the needle valve would fall down.

Owing to the water hammer effect [37], the fluctuations and damping of the compression wave in the common

rail can be significant, giving rise to the pressure oscillation later on. A nozzle opening delay (NOD) time is

defined as the time interval between the start of the current and the start of injection. On the other hand, a nozzle

closure delay (NCD) is defined as the time interval between the end of the current and the end of injection [38]. The

effective injection duration can be calculated as ET + NCD - NOD. The injection current profiles were measured

with a Tektronix A622 current probe (Fig. 1). As shown in Fig. 4b, the NOD time is found to range from 0.30 to 0.40

ms, and it is insensitive to the injection pressure and the injection energizing time. This result is consistent with the

result of Lee [36], which has a smaller nozzle hole diameter.

Figure 5 shows the instantaneous mass flow rate for the rail pressures of 80, 100, 120 and 140 MPa, respectively.

The energizing time has been varied from 0.5 ms to 1.8 ms. The injected fuel mass under each condition is given in

Fig. 6a. A wide range of energizing time is considered to cover the full load and part load operating conditions in

o
internal combustion engines. The fuel is pure diesel. The density is 0.82 g/cm3 @ 20 C and the kinematic viscosity

o
is 3.3 cSt @ 20 C . As seen in Fig. 5, the injection rate profile is varied from triangle shape under short energizing

time to trapezoid shape under long energizing time. Fluctuations of mass flow rate can be seen after injection, Fig.

5. The injection velocity in this period is rather low; the low-momentum fuel spray could deteriorate the

combustion process and produce more NOx and hydrocarbons (HC). For the same injection pressure, the initial

slope of mass flow rate is roughly the same for different energizing times. With a higher injection pressure, the initial

slope of the injection rate profile is higher, since the counter-acting forces are overcome more quickly [39]. The

variation of injection pressure shows rather minor effect on the injection delay. The influence of the injector

energizing time and the injection pressure on the accumulated mass injected can be seen in Fig. 6.
7
The injector energizing time is used to control the injection duration and consequently the injected mass. Fig. 6a

shows a linear correlation between the injected mass and energizing time for the energizing time above 1.1 ms. For

the energizing time below 1.1 ms, a significant kink is observable. The reason is that the peak mass flow rate for a

short energizing time increases with the energizing time, cf. also Fig. 5. The injection rate profile is reduced to a

triangle shape with a single peak mass flow rate, as shown in Fig. 5. In addition to the energizing time, the rail

pressure can also affect the maximal mass flow rate as well as the total injected fuel mass; both quantities increase

with the increased injection pressure. As shown in Fig. 6b, for a short energizing time, the variation of the energizing

time shows more significant effect on the peak mass flow rate than the rail pressure does, owing to its direct impact

on the injection rate profile.

Figure 7 shows the instantaneous mass flow rate of Injector B with the energizing time of 1.4 ms with the rail

pressure varied from 50 MPa to 150 MPa. The data were taken from the experiment of Lee [36]. Similar to that of

injector A, the injection pressure affects the maximum injection rate as well as the slope of the initial and end injection

rate profile. The injection rate profile is similar for a given injection duration and shows a trapezoid shape, owing to

the long energizing time (1.4 ms).

Figure 8 shows the actual injection duration of Injector A and B with the different energizing time, with the rail

pressure varied from 50 MPa to 150 MPa. The injection duration is not equal to energizing time. For the long ET

(greater than 0.4 ms), the injection duration is longer than the energizing time. For the short ET (less than 0.4 ms),

the injection duration is shorter than the ET. With an increasing energizing time, the injection duration is insensitive

to the injection pressure and linear to ET. In most engine operations, the ET usually greater than 0.6 ms, so it is not

suitable for using the ET to represent the injection duration in the simulation of the engine.

8
3. Mathematical modeling of injection rate

3.1. Model formulation

Desantes et al. [40] proposed an injection rate model based on the following injection parameters: start of injection,

slope for boot injection, boot pressure, boot length, peak injection pressure, slope at the end of injection, and injection

time. Following the same idea and based on the injection rate profiles of Injector A and B, an injection rate model for

common rail injection system is sketched in Fig. 9. In the model, the slope at SOI depends on the injection pressure

and the geometry of the injector; the slope at end of injection (EOI) is due to the speed of valve closing.

In order to describe the injection rate profile, the injection process is modeled as five stages, cf. Fig. 9. Stage 1 is

the start of injection (0~ t1 ), during which the needle valve begins to open and the injection velocity increase rapidly.

Stage 2 is the quasi-steady injection period ( t1 ~ t2 ), during which the needle valve is nearly fully open and valve

lifts slowly with a moderate increase of the injection velocity. Stage 3 is the peak injection rate period ( t2 ~ topen ),

during which the needle valve is completely open and the injection velocity is at its maximum. Stage 4 is the valve

closing period ( t4 ), during which the needle valve falls back into valve sit, causing the injection rate to decrease.

Stage 5 is the end of injection ( t5 ); at the end of this period the needle valve is completely close and the injection

velocity decreases to zero. In this five-stage model, the flow rate fluctuation after the end of injection, as shown in

Fig. 5, is ignored. Note that for the very short energizing time, topen is less than t1 and the injection stages 2 and

3 do not appear. For a moderately long energizing time with topen less than t2 the injection stage 3 does not appear.

In the injection rate model, key model parameters are the time duration of the different stages, t1 up to t5 , topen

and tclose , as well as the maximal injection rate. Determination of the injection profiles in various stages and the time

duration of the stages will be described below. First, the maximal injection rate is a parameter to be determined. A

commonly used approach (cf. Refs. [37], [41] and [42]) to calculate the maximum injection rate is based on the nozzle

hole area and the injection velocity:

9
G = nCd ρ f A0uth (1)

where G is the maximal mass flow rate, n is the number of the injector holes, A0 is the geometrical orifice

outlet area , Cd is the nozzle discharge coefficient and uth is the theoretical velocity calculated from the

Bernoulli's equation for viscous flows:

2∆P (2)
uth =
ρf

where ∆P= Prail − Pb is the pressure difference between the injection pressure and ambient pressure, ρ f is the

fuel density. In realistic injectors, the peak injection velocity cannot reach uth due to friction loss and cavitation

inside the injector nozzle. The discharge coefficient Cd is used to account for the loss of injection velocity. The

nozzle hole discharge coefficient is extensively studied in literature [43]; however, there is no consensus in the

suggested value of this coefficient. Yu et al [43]. reported a discharge coefficient of the values of 0.64-0.72 for

cavitation numbers in the range of 1.0-1.3. Nurick et al. [44] presented the discharge coefficients of solenoid and

piezoelectric nozzles around 0.65-0.78. Salvador et al. [45] reported a discharge coefficient between 0.76 and 0.86,

which increased with the Reynolds number. As high as 0.8-0.88 of discharge coefficient was reported by Payri et al.

[46], when the Reynolds number was in the range of 2000-120000.

The inconsistent discharge coefficient in literature poses a challenge of accurate modeling of the injection rate.

With the present measurements of the injection rate we can determine the discharge coefficient for the present

injectors. According to Eq. (1), the discharge coefficient can be calculated [47]:

G G
=Cd = (3)
n ρ f A0uth nA0 2 ρ f  p

where G is the maximal mass flow rate of the liquid fuel obtained in the experiment. The Experiment data of injector

B were used to calculate the discharge coefficient and the results are shown in Table 3, where the mass flow rate for

one hole has been shown. The geometrical parameters of injector B can be found in Table 2.

10
In Refs. [45, 48, 49] the discharge coefficient is expressed as a function of Reynolds number. The Reynolds number

is defined as:

uth D0 (4)
Re =
νf

where ν f is the liquid fuel kinematic viscosity and the Do is the nozzle exit diameter. The discharge coefficient

defined in Eq. (3) for the injector B is given in Table 3 along with the Reynolds number. Fig. 10 shows the discharge

coefficient against the Reynolds number for both injector A and B. It is found that the discharge coefficient increases

with the Reynolds number. The relationship between discharge coefficient and Reynolds number, Re, can be fitted

into a quadratic function:


2
 10000   10000  (5)
−0.3128 
Cd =  + 0.1867   + 0.8057
 Re   Re 

In the engine operation, the injection pressure usually can be equal to the rail pressure. The injection duration and

injection mass flow rate can be determined from the injection pressure and the injected total fuel mass. For the five-

stage injection rate model sketched in Fig. 9, the profile of the first three stages can be approximated with the

following polynomials:

G1 (t ) = a1t 3 + b1t 2 + c1t 0 ≤ t ≤ t1 (6)

G2 (t ) = a2t 2 + b2t + c2 t1 ≤ t ≤ t2 (7)

G3 (t ) = Gmax t2 ≤ t ≤ topen (8)

From the experimental data of injector B it is found that t1 = 0.4 ms, t2 = 1.2 ms. topen is the time duration of

valve opening. It is typically unknown in the injection experiments. This parameter is to be determined in an iterative

procedure to be discussed later. The values of a2 and b2 are respectively -0.811 and 2.367, obtained by fitting the

experimental data. It appears that these constants are independent of the injection pressure. In the present paper, the

11
injection rate model parameters have the unit that is based on mg and ms. Figure 11b shows a comparison between

the model and the experimental data for the second stage of the injection process. The coefficient, c2 , is,

c2= Gmax − a2 ×1.22 − b2 ×1.2= Gmax − 1.672 (9)

where Gmax is the maximum of injection rate at a given injection pressure, which is calculated using Eq. (10). The

discharge coefficient Cd used in Eq. (10) is calculated by Eq. (5).

Gmax = Cd ρ f A0uth (10)

Figure 11a shows the injection rate profile for the first stage of injection, 0~0.4 ms. It is clear that the model can

replicate the experimental data with a cubic function. The slope of the injection rate profile at the start of injection is

equal to c1 , which shows the rate of injection velocity acceleration and has a linear correlation with the force on the

injector hole. It can be calculated as follows:

=c1 16.22 ∆PA 0 (11)

The coefficients a1 and b1 can be calculated based on G1 ( t ) and G2 ( t ) by matching the slope and value

at t1 . From Eqs. (12) ~ (16), it appears that

G1' (t1 ) = 3a1t12 + 2b1t1 + c1 (12)

'
G2= (t1 ) 2a2t1 + b2 (13)

G1' (t1 ) = G2 ' (t1 ) (14)

G1' ( t1 ) t1 − 2G1 ( t1 ) + c1t1


a1 = (15)
t13

G1' ( t1 ) − 3a1t12 − c1
b1 = (16)
2t1

In stage 4 the needle valve starts to close, and the injection velocity begins to decrease from its peak value. Because

of the water hammer effect, the needle valve cannot close immediately and the flow rate decreases gradually.
12
According to the experiment, the duration of this stage is very short. The period of stage 4 is modeled as t4 =k1t peak ,

where k1 is

k1 =0.1t peak 2 − 0.32t peak + 0.42 (17)

where t peak is the time at which the peak value of the injection rate is achieved.

t peak = min ( t2 ,topen ) (18)

The mass flow rate profile in stage 4 can be fitted into a quadratic function:

G4 (t ) a4 ( t - topen ) + G ( topen )
2
= topen ≤ t ≤ topen + t4 (19)

The injection stage 5 can be fitted into a linear function. The mass flow rate decreases to zero in a very short time,

t5 . The period of stage 5 is assumed to , where k2 is calculated from Eq. (21). tid is the total injection

duration tid = topen + ( k1 + k2 ) t peak .

G5 (t ) = a5 ( t − topen − t4 ) + c5 topen + t4 ≤ t ≤ tid (20)

k2 = −0.1t peak + 0.16 (21)

The coefficients, a4 , a5 and c5 can be calculated by matching the same slope and value of G4 ( t ) and

G5 ( t ) at topen + t4 . From Eqs. (17) ~ (21), it appears that

G ( topen )
a4 = - (22)
t4 ( 2t5 + t4 )

a5 = 2a4t4 (23)

=c5 G ( topen ) + a4t42 (24)

Figure 12 shows the flow chart of the computational procedure with the new injection model. In step 1, six input

parameters are specified, namely, the fuel density, the kinematic viscosity, the injection pressure, the total injected

fuel mass, the nozzle hole diameter, and the number of the injector holes. In step 2, the discharge coefficient Cd

and the maximal injection rate Gmax at a given operation condition are calculated using Eqs. (5) and (10). In step 3,

the model coefficients appeared in Eqs. (6), (7), and (8) are calculated using Eqs. (9) ~ (16). In step 4, the valve open
13
time topen is presumed and the model coefficients appeared in Eqs. (19), (20) are calculated using Eqs. (22) ~ (24).

In step 5, the total injected mass is computed by integration of the injection profile G ( t ) from 0 to tid . The

integration is compared with the total fuel mass injected, and if the difference is greater than a tolerable error (e.g.

error=0.001) a new value of topen is set. The procedure is continued until the error of the injected mass is lower than

the tolerable value. If the topen is assumed known (e.g. to be equal to the energizing time in the case of long

energizing time, as shown in Fig. 13), the total fuel mass can also be calculated based on the energizing time and the

injection pressure.

3.2. Validation for injection rate model

The injection rate model is validated by comparing the model predictions with the injection rate profiles obtained

in different experiments. In all simulations shown in this section, the time duration of the first two stages of injection

has been kept constant, with t1 = 0.4 ms, t2 = 1.2 ms, and the values of a2 and b2 kept as -0.811 and 2.367.

These parameters appear to be universal and independent of the injection pressure or energizing time. The values of

tolerable error set as 0.001mg, which is less than 0.01% of the total injected mass for most cases. The time step  t

is set to be 0.0001 ms.

o
The experiments of Injector B [36] are considered first. The fuel density is 0.848 g/cm3 @ 20 C and the

kinematic viscosity is 3.3 cSt @ 20 oC . Figure 14 shows the temporal evolution of the mass flow rate from the

experiments and model predictions at different rail pressures. The results indicate that the new injection rate model

is able to reasonably reproduce the instantaneous mass flow rate under different injection pressures and energizing

times.

14
It is clear that the new model gives satisfactory results for Injector B. To further evaluate the model, the experiments

with Injector A are considered. As shown in Fig. 15, the results predicted by the model are in reasonably good

agreement with the experiments for the long energizing times. The first stage injection profile appears to be in good

agreement with the experiment for all injection conditions, indicating the validity of the assumption of universal

constants for model parameters t1 , t2 , a2 and b2 . For short energizing time cases, e.g. less than 0.7 ms, the error

in the peak mass flow rate is relatively large, which can be attributed to that, for the short energizing time, the needle

value open and close quickly, causing a more significant fluctuation of pressure in the injector. In the experiments of

injector A, due to the rapid pressure oscillation there is rather large uncertainty in the mass flow rate.

The experiment of Seykens et al. [50] is also considered for validation of the injection model; the case is referred

hereafter as Injector C. Injector C has eight holes with a hole diameter of 0.184 mm. Diesel and Rapeseed Methyl

Ester (RME) were used in the experiments. The physical properties of diesel and RME are given in Table 4. The

available experimental data were measured at an energizing time 3.8 ms with a single injection. Four rail pressures

of 80, 100, 120 and 140 MPa were chosen for the validation of the present model. As seen in Fig. 16, the predicted

mass flow rates of both diesel and RME are in satisfactory agreement with the experiments.

The injection model is also validated against the experimental data of Agarwal et al. [51] (hereafter referred to as

Injection D). The total injected fuel mass was specified at 12 mg and the injection pressures were varied from 500

o o
bar to 1000 bar. The density of the fuel is 0.831 g/cm3 @ 40 C and the kinematic viscosity is 2.78 cSt @ 40 C .

As shown in Fig. 17, with the increased fuel injection pressure, injection duration decreases and the peak injection

rate increases. The new model replicates the effect of injection pressure on injection duration and peak injection rate

very well.

15
4. Application of Injection Rate Model to Diesel Engines

The model is integrated into the KIVA-3V code to simulate a 5.3-liter four-cylinder diesel engine in order to

evaluate the capabilities of the new model for prediction of the effects of injection pressure and energizing time on

the injection mass flow rate, and on the combustion and emission process. The results are compared with the engine

pressure trace and heat release rate profiles, and emission data. Experimental study of the engine has been performed

at Shanghai Jiao Tong University. The in-cylinder gas pressure was measured by a pressure transducer (Kistler model

6125B). The charge output was converted to an amplified voltage using an amplifier (Kistler model 5015A). The heat

release rate and in-cylinder temperature were calculated using the D2T combustion analyzer. The total unburned

hydrocarbon (HC) and NOx emissions were measured using a heated flame ionization detector (HFID CAI 600).

More detailed information about the experiment can be found in Ref. [52].

4.1. Numerical Models

The simulation was performed with the improved turbulent combustion sub-models in the KIVA-3V code. The

turbulence was modeled using the generalized renormalization Group (GRNG) κ-ε turbulence model with adjusted

model coefficients for variable-density flows, which can perform better with compressing/expanding flows in a

diesel engine [53]. The Kelvin-Helmholtz (KH) instability model was used to predict the primary breakup and the

Rayleigh-Taylor (RT) accelerative instability model was used to predict the secondary breakup [28]. The coefficients

of the KH-RT hybrid model were modified based on the experiment. The spray collision model [54] with enhanced

grid independence and the spray/wall interaction model [55] were adopted. The heat-transfer model was based on the

improved model of Han et al. [56] by considering the turbulent Prandtl number. n-Heptane is chosen as the surrogate

of diesel fuel due to their similar chemical properties [57]. A skeletal mechanism of n-heptane developed by Chang

et al. [58] was used to model the chemical reactions. The mechanism consists of 176 elementary reactions

involving 56 species. In addition, the extended Zeldovich mechanism [59] and phenomenological soot model
16
[60] were coupled with the chemical mechanism to simulate the formulation of NOx and soot. The new injection

model was integrated into the KIVA-3V code, which was applied for the prediction of injection rate profile and

injection duration.

The simulation starts from IVC (-145.5 oCA ATDC) and ends at EVO (112 oCA ATDC). The emission data of HC

and NOx are taken at EVO. For initial conditions, the charge temperature, pressure, and species densities were

assumed to be uniform in an entire combustion chamber. The initial flow inside a cylinder was assumed to be a solid-

body rotational flow with a swirl ratio of 1.9. The specification of the diesel engine and operating conditions are

given in Table 5.

4.2. Sensitivity to grid density

The computational mesh is made up of a sector of the combustion chamber, which corresponds to one spray plume.

Two different computational grids (hereafter referred to as coarse mesh and fine mesh) were constructed to study the

effects of grid density on the model results for a baseline injection condition (IMEP=6 bar, rpm of 1500) with respect

to the in-cylinder pressure profile and heat release rate, as shown in Fig. 18. The total numbers of cells of the coarse

mesh and fine mesh are respectively 15287 and 31317 at bottom dead center (BDC) and 6891 and 10355 at top dead

center (TDC).

Figure 19 shows that the two meshes yield similar in-cylinder pressure and heat release rate, which indicates that

the results are grid insensitive with the coarse mesh. Thus, the remaining results are obtained using the coarse mesh.

5. Results and discussion

First, the engine speed of 1500 rpm case is studied. In order to validate the accuracy of the new injection model in

predicting the injection velocity profile under different injection pressures and fuel masses, the conditions of IMEP

of 4-7 bar were investigated, which covers the injection pressure range from 80 to 120 MPa, corresponding to part-

17
to-high load conditions. The injection duration and injection rate are calculated using the new model under these

conditions, as shown in Table 6. It should be noted that all other engine parameters such as the injection timing, initial

temperature, and initial pressure given in Table 5 are kept unchanged.

Figure 20a shows the injection rate profile for the different injected fuel mass at 100 MPa injection pressure. The

shape of the profile changes from triangle to trapezoid with increasing injected mass (and injection duration). For the

cases with the same injected mass, the peak mass flow rate and injection duration vary with the injection pressure,

Fig. 20b. Higher injection pressure produces the higher mass flow rate and as such a shorter injection duration is

permitted to maintain a constant injected fuel mass.

Figure 21 shows the in-cylinder pressure and heat release rate at different crank angles for the various injection

pressures and loads studied, along with the emissions of NOx and HC. The predicted in-cylinder pressure and heat

release rate show reasonable agreement with the experimental dates for all cases, in particular for the onset of

combustion and combustion duration. The over-prediction of the pressure-rise rate and the heat release rate is a result

of the used RANS turbulent combustion model, which tends to smoothen the temperature and concentration fields,

and as such a more rapid combustion rate is often predicted.

HC emissions are serious problem at light loads for diesel engines; this is due to the low combustion temperature

for the fuel oxidation. With higher load, the increased combustion temperature leads to lower HC emissions and

higher NOx emissions. This trend is correctly predicted in the simulations. NOx emissions are over-predicted for all

the injection pressures and the predicted HC emission at low load is higher than the experimental data. The

quantitative difference of HC and NOx emissions between the numerical simulation and experiment is attributed to

the relatively simple model employed in this study in compromise of the computational time. Overall, the HC and

NOx variation with increased load can be qualitatively predicted with the new injection rate model.

18
Next, the combustion process in the Diesel engine at two other engine speeds, 1200 and 1800 rpm was investigated

under different loads with the constant injection pressures of 120 MPa. As given in Table 5, the NOD time is 0.32

ms. The injection timing in the cases is -7 CA ATDC; as such, the start of the injection is -4.696 and -3.544 CA ATDC

for 1200 and 1800 r/min.

Figures 22 show the variations of in-cylinder pressure and heat release rate against the crank angle at different

engine loads and engine speeds. The predicted in-cylinder pressure and heat release rate profiles show reasonably

good agreement with the experimental data for all cases. The simulation can accurately predict the ignition delay and

CA50 (the crank angle at which 50% of heat is released) at different engine speeds. At the same load, the in-cylinder

peak pressure and heat release rate decrease with the engine speed. The time left for fuel oxidation in millisecond

(ms) decreases with engine speed, so the combustion efficiency decreases accordingly, resulting in a higher HC

emission (also high CO emission; for brevity, the result of CO is not shown in the figure) and a lower NOx emission.

It is observed that the trend of NOx and HC emissions observed in the experiments are well captured in the numerical

simulations. Overall, the numerical simulation with the new injection model is capable of replicating the experimental

data, including in-cylinder pressure, heat release rate, and emission characteristics for direct inject combustion.

Figure 23 shows the evolution of equivalence ratio and spray in the cylinder at different injection pressures of 80,

100 and 120 MPa, under the engine speed of 1500 rpm. The parcels represent the liquid fuel droplets. The vertical

dashed line indicates the liquid penetration length. It is observed that the liquid penetration length increases with the

high rail pressure [61] (cf. -2 CA ATDC). Higher injection pressure can provide higher kinetic energy for the spray

droplets and the interaction between droplets and ambient gas is enhanced. The size of the injected fuel parcels is

smaller with a higher injection pressure [62]; thus, the mixing process is faster with increasing injection pressure.

The maximum local equivalence ratio at 5 CA ATDC is 5.09, 4.08, and 3.6 under the injection pressure of 80, 100,

and 120 MPa, respectively. Hence, high injection pressure results in faster spray penetration and spreading, and more

19
homogeneous distribution of local equivalence ratio [63]. The equivalence ratio field is affected by the flow swirl,

which is along the clockwise direction. It is seen that the large parcels are not affected by the swirl but the smaller

ones tend to follow the flow more closely.

The in-cylinder temperature distribution for different injection pressures is shown in Fig. 24. The results show that

the ignition takes place first around the wall of the piston region (5 oCA ATDC) and is widely distributed in space.

It is observed that the temperature distribution is more homogeneous for the high injection pressure case (120 MPa).

Fig. 23 has demonstrated that a higher injection pressure leads to a faster liquid spray penetration and faster mixing

of the vapor fuel with the ambient air. Thus, the combustion process takes place in a wider space throughout the

whole combustion chamber, and the temperature field is more uniform. The ignition takes place at further

downstream and the flame spreads more quickly into the squish region. This explains the increased NOx emission

with increasing injection pressure as already shown earlier in Fig. 21.

6. Conclusion

The liquid fuel injection process plays a key role in the fuel/air mixing, and the combustion and emission processes

in direct injection engines, Injection rate and injection duration as the key input parameters are required in the

numerical simulation of the spray development and combustion process in the diesel engine cylinder. Although the

injected fuel mass per cycle can be measured easily in the experiment, the fuel injection rate profile is hardly measured.

Thus, very rare experimental data are available for modeling the injection rate, which is usually simply assumed to

be of square, trapezoid, or Gaussian shaped profile. Accurate representation of the injection rate for a given operating

condition (rail pressure and fuel mass) is therefore of great importance. In this work, the effects of injection pressure

and energizing time on the injection rate profile of a common rail injection system fueled with diesel are investigated

experimentally and numerically. Based on the present experimental data and experimental data reported in the

20
literature, a new injection rate model for the common rail fuel injection systems is developed. The new injection

model is integrated into a CFD code to study the combustion process and emissions formation in a diesel engine. The

following conclusions are drawn.

• The variation of the injection pressure shows little effect on the nozzle opening delay (NOD) time. The

experimental results showed that the delay time ranges from 0.30 to 0.40 ms. The experiment indicates

that there is no direct relationship between the nozzle opening time and the energizing time registered

in the electronic control unit (ECU). Subsequently, the relationship between energizing time and the

total injected fuel mass is not linear. A significant kink is observable for the energizing time below 1.1

ms. With an increasing energizing time, the shape of the injection rate profile changes from triangle

for short energizing time to trapezoid for long energizing time. The peak mass flow rate for the short

energizing time is lower than the peak mass flow rate for the long energizing time.

• The injection process can be characterized as five stages. The first stage is a fast valve opening stage,

during which the injector nozzle valve opens quickly after a short NOD time, and the injection rate

increases rapidly with time. It is found that this stage is insensitive to the injector nozzle diameter, the

injection pressure, and injection duration. The second stage is the valve slow opening stage, during

which the needle lift slows down until the valve is fully open. For the present injection experiment

under sufficiently long injection duration, this stage appears to start around 0.4 ms after the start of

injection (ASOI) and end around 1.2 ms ASOI, regardless of the injection pressure and injection

duration. The same conclusion is found when examining the experimental data reported in the literature.

• A new injection rate model is proposed based on the present experimental data and other experimental

data reported in the literature. The model is based on the five-stage injection process, with a universal

NOD time, and universal first stage injection model parameters, a peak mass flow rate modeled based

21
on an empirical nozzle hole discharge coefficient, and experimentally determined total injected fuel

mass and energizing time. The new injection rate model is capable of accurately predicting the injection

rate profiles and injection duration for varying injection pressure and energizing time under various

experimental conditions and injector configurations.

• The nozzle hole discharge coefficient is a key model parameter for the injection rate model. The

discharge coefficient strongly influences the maximum injection mass flow rate. An empirical

expression of the discharge coefficient as a function of the Reynolds number is proposed and the

expression is consistent with the present experimental data as well as data from the literature. The new

injection rate model successfully replicated the injection process in a diesel engine, which is proved by

the reasonable predictions of the in-cylinder pressure, ignition timing and combustion duration, as well

as the emissions of HC and NOx. The model correctly captured the effect of injection pressure and

injection duration on the spray atomization, vapor fuel/air mixing, and the combustion process.

• According to the current experiment data and the date in the literature, the parameters in the first two

stages of injection, including t1 , t 2 , a2 , and b2 , appear to be universal and independent of the

injector. However, more experiment validations are required to verify the assumption. Since the

influences of cavitation and other factors (e.g. turbulent flows) are not explicitly modeled, further

improvement of the discharge coefficient formulation is necessary. The effect of the fuel properties

and injector geometry on the injection rate should also be considered to extend the injection rate model.

Acknowledgements

This work was supported by the Natural Science Foundation of China (Grant No. 51425602) and High Technology

Ship Research Program-Marine Low Speed Engine (Phase I). The authors gratefully acknowledge the financial

22
support from the China Scholarship Council (No. 201706230040) for the visit of the first author to Lund University,

Sweden.

Abbreviation

CRIS: Common Rail Injection System

HCCI: Homogeneous charge compression ignition

PCCI: Premixed charge compression ignition

PPC: Partially premixed combustion

RCCI: Reactivity controlled compression ignition

EGR: Exhaust gas recirculation

DI: Direct ignition

MK: Modulated Kinetics

LTC: low temperature combustion

TAB: Taylor Analogy Breakup

CFD: Computational Fluid Dynamics

LISA: Linearized Instability Sheet Atomization

KH-RT: Kelvin Helmholtz -Rayleigh Taylor

ECU: Electronic control unit

ROI: Rate of injection

ET: Energizing time

NOD: Nozzle opening delay

NCD: Nozzle closure delay

23
NOx: Nitrogen oxides

HC: Hydrocarbons

G: Mass flow rate

n: Number of the injector holes

A0 : Geometrical orifice outlet area

Cd : The nozzle discharge coefficient

uth : The theoretical injection velocity

Prail : Rail pressure

Pback : Back pressure

P : pressure drop

ρf : Fuel density

Re: Reynolds number

νf : Fuel kinematic viscosity

Do : The nozzle exit diameter

Gmax : Maximum of injection rate at specified injection pressure

topen : The period that the needle valve keeps open

t peak : The minimal time at which the peak value of injection rate is achieved

ASOI: After the start of injection

RME: Rapeseed Methyl Ester

DI: Direct injection

GRNG: Generalized renormalization Group

IMEP: Indicated mean effective pressure

24
BDC: Bottom dead center

TDC: Top dead center

ATDC: After top dead center

CA: Crank angle

CA50: Crank angle at which 50% of heat is released

Subscripts

f : liquid fuel

peak: the maximum rate of injection profile

open: the needle valve keep open

rail: common rail

max: maximum value

th: theoretical value

id: injection duration

1: the first injection part

2: the second injection part

3: the third injection part

4: the fourth injection part

5: the fifth injection part

25
Reference

[1] Lu X, Han D, Huang Z. Fuel design and management for the control of advanced compression-ignition combustion

modes. Prog Energy Combust Sci 2011;37:741-83.

[2] Kiplimo R, Tomita E, Kawahara N, Yokobe S. Effects of spray impingement, injection parameters, and EGR on the

combustion and emission characteristics of a PCCI diesel engine. Appl Therm Eng 2012;37:165-75.

[3] Noehre C, Andersson M, Johansson B, Hultqvist A. Characterization of partially premixed combustion. SAE Technical

Paper, 2006-01-3412, 2006.

[4] Benajes J, Molina S, García A, Monsalve-Serrano J, Durrett R. Performance and engine-out emissions evaluation of

the double injection strategy applied to the gasoline partially premixed compression ignition spark assisted combustion

concept. Appl Energy 2014;134:90-101.

[5] Kokjohn SL, Hanson RM, Splitter D, Reitz R. Fuel reactivity controlled compression ignition (RCCI): a pathway to

controlled high-efficiency clean combustion. Int J Engine Res 2011;12:209-26.

[6] Benajes J, Molina S, García A, Monsalve-Serrano J. Effects of low reactivity fuel characteristics and blending ratio on

low load RCCI (reactivity controlled compression ignition) performance and emissions in a heavy-duty diesel engine.

Energy 2015;90:1261-71.

[7] Kimura S, Aoki O, Kitahara Y, Aiyoshizawa E. Ultra-clean combustion technology combining a low-temperature and

premixed combustion concept for meeting future emission standards. SAE Technical Paper, 2001-01-0200, 2001.

[8] Kimura S, Aoki O, Ogawa H, Muranaka S, Enomoto Y. New combustion concept for ultra-clean and high-efficiency

small DI diesel engines. SAE Technical Paper, 1999-01-3681,1999.

[9] Zhang B, Jiaqiang E, Gong J, Yuan W, Zuo W, Li Y, et al. Multidisciplinary design optimization of the diesel particulate

filter in the composite regeneration process. Appl Energy 2016;181:14-28.

26
[10] Deng Y, Liu H, Zhao X, Jiaqiang E, Chen J. Effects of cold start control strategy on cold start performance of the

diesel engine based on a comprehensive preheat diesel engine model. Appl Energy 2018;210:279-87.

[11] Agarwal AK, Srivastava DK, Dhar A, Maurya RK, Shukla PC, Singh AP. Effect of fuel injection timing and pressure

on combustion, emissions and performance characteristics of a single cylinder diesel engine. Fuel 2013;111:374-83.

[12] Wang B, Jiang Y, Hutchins P, Badawy T, Xu H, Zhang X, et al. Numerical analysis of deposit effect on nozzle flow

and spray characteristics of GDI injectors. Appl Energy 2017;204:1215-24.

[13] Moon S, Huang W, Li Z, Wang J. End-of-injection fuel dribble of multi-hole diesel injector: Comprehensive

investigation of phenomenon and discussion on control strategy. Appl Energy 2016;179:7-16.

[14] Gentz G, Gholamisheeri M, Toulson E. A study of a turbulent jet ignition system fueled with iso-octane: Pressure trace

analysis and combustion visualization. Appl Energy 2017;189:385-94.

[15] Lefebvre AH, McDonell VG. Atomization and sprays: CRC press; 2017.

[16] Kashdan JT, Anselmi P, Walter B. Advanced injection strategies for controlling low-temperature diesel combustion

and emissions. SAE Paper 2009-01-1962, 2009.

[17] Tay KL, Yang W, Zhao F, Yu W, Mohan B. Effects of triangular and ramp injection rate-shapes on the performance

and emissions of a kerosene-diesel fueled direct injection compression ignition engine: A numerical study. Appl Therm

Eng 2017;110:1401-10.

[18] Boggavarapu P, Singh S. Computational study of injection rate-shaping for emissions control in diesel engines. SAE

Technical Paper, 2011-26-0081, 2011.

[19] Mohan B, Yang W, Yu W, Tay KL, Chou SK. Numerical investigation on the effects of injection rate shaping on

combustion and emission characteristics of biodiesel fueled CI engine. Appl Energy 2015;160:737-45.

[20] Rottmann M, Menne C, Pischinger S, Luckhchoura V, Peters N. Injection Rate Shaping Investigations on a Small–

Bore DI Diesel Engine. SAE Technical Paper, 2009-01-0850, 2009.

27
[21] Tanabe K, Kohketsu S, Nakayama S. Effect of fuel injection rate control on reduction of emissions and fuel

consumption in a heavy duty DI diesel engine. SAE Technical Paper, 2005-01-0907, 2005.

[22] Wakisaka Y, Azetsu A. Effect of fuel injection rate shaping and injection pressure on intermittent spray combustion.

SAE Technical Paper, 2000-01-2793, 2000.

[23] Macian V, Payri R, Ruiz S, Bardi M, Plazas AH. Experimental study of the relationship between injection rate shape

and Diesel ignition using a novel piezo-actuated direct-acting injector. Appl Energy 2014;118:100-13.

[24] Agarwal AK, Som S, Shukla PC, Goyal H, Longman D. In-nozzle flow and spray characteristics for mineral diesel,

Karanja, and Jatropha biodiesels. Appl Energy 2015;156:138-48.

[25] Reitz RD, Bracco F. Mechanisms of breakup of round liquid jets. Encyclopedia of fluid mechanics 1986;3:233-49.

[26] Senecal P, Schmidt DP, Nouar I, Rutland CJ, Reitz RD, Corradini M. Modeling high-speed viscous liquid sheet

atomization. Int J Multiphase Flow 1999;25:1073-97.

[27] O'Rourke PJ, Amsden AA. The TAB method for numerical calculation of spray droplet breakup. SAE Technical Paper,

872089, 1987.

[28] Patterson MA. Modeling the effects of fuel injection characteristics on diesel combustion and emissions. PhD Thesis,

University of Wisconsin-Madison; 1997.

[29] Richards K, Senecal P, Pomraning E. CONVERGE 2.1. 0 Theory Manual, Convergent Science. Inc., Middleton, WI

2013.

[30] Jasak H. OpenFOAM: open source CFD in research and industry. International Journal of Naval Architecture and

Ocean Engineering 2009;1:89-94.

[31] Anthony Amsden L. A Block-Structured KIVA Program for Engines with Vertical or Canted Valves. Los Alamos

National Laboratory; 1999.

28
[32] Čaika V, Sampl P, Tatschl R, Krammer J, Greif D. Coupled 1D-3D simulation of common rail injector flow using AVL

HYDSIM and FIRE. SAE Technical Paper, 2009-24-0029, 2009.

[33] Bianchi GM, Falfari S, Parotto M, Osbat G. Advanced modeling of common rail injector dynamics and comparison

with experiments. SAE Technical Paper, 2003-01-0006, 2003.

[34] Imagine S. AMESim and ‘Common Rail’Type Injection Systems. Technical Bulletin n 2000;110.

[35] Han D, Zhai J, Duan Y, Ju D, Lin H, Huang Z. Macroscopic and microscopic spray characteristics of fatty acid esters

on a common rail injection system. Fuel 2017;203:370-9.

[36] Lee S-S. Investigation of two low emissions strategies for diesel engines: Premixed charge compression ignition (PCCI)

and stoichiometric combustion. PhD Thesis, University of Wisconsin-Madison; 2006.

[37] Seykens X, Somers L, Baert R. Detailed modeling of common rail fuel injection process. Journal of Middle European

Construction and Design of Cars 2005;3:30.

[38] Ferrari A, Mittica A. Response of different injector typologies to dwell time variations and a hydraulic analysis of

closely-coupled and continuous rate shaping injection schedules. Appl Energy 2016;169:899-911.

[39] Wang Z, Ding H, Ma X, Xu H, Wyszynski ML. Ultra-high speed imaging study of the diesel spray close to the injector

tip at the initial opening stage with single injection. Appl Energy 2016;165:335-44.

[40] Desantes JM, Benajes J, Molina S, González CA. The modification of the fuel injection rate in heavy-duty diesel

engines. Part 1: Effects on engine performance and emissions. Appl Therm Eng 2004;24:2701-14.

[41] Poetsch C. Crank-Angle Resolved Modeling of Fuel Injection and Mixing Controlled Combustion for Real-Time

Application In Steady-State and Transient Operation. No. 2014-01-1095. SAE Technical Paper Series, 2014.

[42] Turner MR, Sazhin SS, Healey JJ, Crua C, Martynov SB. A breakup model for transient Diesel fuel sprays. Fuel

2012;97:288-305.

29
[43] Yu W, Yang W, Zhao F. Investigation of internal nozzle flow, spray and combustion characteristics fueled with diesel,

gasoline and wide distillation fuel (WDF) based on a piezoelectric injector and a direct injection compression ignition

engine. Appl Therm Eng 2017;114:905-20.

[44] Payri R, Salvador FJ, Gimeno J, De la Morena J. Influence of injector technology on injection and combustion

development – Part 1: Hydraulic characterization. Appl Energy 2011;88:1068-74.

[45] Salvador FJ, Lopez JJ, De la Morena J, Crialesi-Esposito M. Experimental investigation of the effect of orifices

inclination angle in multihole diesel injector nozzles. Part 1 – Hydraulic performance. Fuel 2018;213:207-14.

[46] Payri R, García A, Domenech V, Durrett R, Plazas AH. An experimental study of gasoline effects on injection rate,

momentum flux and spray characteristics using a common rail diesel injection system. Fuel 2012;97:390-9.

[47] Emberson DR, Ihracska B, Imran S, Diez A, Lancaster M, Korakianitis T. Hydraulic characterization of Diesel and

water emulsions using momentum flux. Fuel 2015;162:23-33.

[48] Som S, Aggarwal S, El-Hannouny E, Longman D. Investigation of nozzle flow and cavitation characteristics in a

diesel injector. J Eng Gas Turbines Power-Trans ASME, 2010;132:042802.

[49] Boudy F, Seers P. Impact of physical properties of biodiesel on the injection process in a common-rail direct injection

system. Energy Convers Manage 2009;50:2905-12.

[50] Seykens X, Somers L, Baert R. Modelling of common rail fuel injection system and influence of fluid properties on

injection process. Proceedings of VAFSEP 2004:6-9.

[51] Agarwal AK, Dhar A, Gupta JG, Kim WI, Choi K, Lee CS, et al. Effect of fuel injection pressure and injection timing

of Karanja biodiesel blends on fuel spray, engine performance, emissions and combustion characteristics. Energy Convers

Manage 2015;91:302-14.

30
[52] Qian Y, Zhang Y, Wang X, Lu X. Experimental investigation of the combustion characteristics and the emission

characteristics of biogas–diesel dual fuel in a common-rail diesel engine. Proc Inst Mech Eng Pt D: J Automobile Eng

2017:0954407017691398.

[53] Wang B-L, Lee C-W, Reitz RD, Miles PC, Han Z. A generalized renormalization group turbulence model and its

application to a light-duty diesel engine operating in a low-temperature combustion regime. Int J Eng Res 2013;14:279-92.

[54] Nordin P. Complex chemistry modeling of Diesel spray combustion. PhD thesis, Chalmers University of Technology;

2001.

[55] Han Z, Xu Z, Trigui N. Spray/wall interaction models for multidimensional engine simulation. Int J Eng Res

2000;1:127-46.

[56] Han Z, Reitz RD. A temperature wall function formulation for variable-density turbulent flows with application to

engine convective heat transfer modeling. Int J Heat Mass Transfer 1997;40:613-25.

[57] Saisirirat P, Togbé C, Chanchaona S, Foucher F, Mounaim-Rousselle C, Dagaut P. Auto-ignition and combustion

characteristics in HCCI and JSR using 1-butanol/n-heptane and ethanol/n-heptane blends. Proc Combust Inst

2011;33:3007-14.

[58] Chang Y, Jia M, Li Y, Liu Y, Xie M, Wang H, et al. Development of a skeletal mechanism for diesel surrogate fuel by

using a decoupling methodology. Combust Flame 2015;162:3785-802.

[59] Williams A, Pourkashanian M, Bysh P, Norman J. Modelling of coal combustion in low-NOx pf flames. Fuel

1994;73:1006-19.

[60] Jia M, Peng Z, Xie M. Numerical investigation of soot reduction potentials with diesel homogeneous charge

compression ignition combustion by an improved phenomenological soot model. Proc Inst Mech Eng Pt D: J Automobile

Eng 2009;223:395-412.

31
[61] Agarwal AK, Dhar A, Gupta JG, Kim WI, Lee CS, Park S. Effect of fuel injection pressure and injection timing on

spray characteristics and particulate size–number distribution in a biodiesel fuelled common rail direct injection diesel

engine. Appl Energy 2014;130:212-21.

[62] Jiang C, Xu H, Srivastava D, Ma X, Dearn K, Cracknell R, et al. Effect of fuel injector deposit on spray characteristics,

gaseous emissions and particulate matter in a gasoline direct injection engine. Appl Energy 2017;203:390-402.

[63] Jiaqiang E, Pham M, Deng Y, Nguyen T, Duy V, Le D, et al. Effects of injection timing and injection pressure on

performance and exhaust emissions of a common rail diesel engine fueled by various concentrations of fish-oil biodiesel

blends. Energy 2018;149:979-89.

32
Figure Captions:

Fig. 1 Sketch of injection rate measurement system

Fig. 2 The injection rate experimental system

Fig. 3 Temporal evolution of the instantaneous energizing current and ROI signals for 20 injections (a), and mean

averaged mass flow rate from the 20 injections (b) for Injector A.

Fig. 4 Temporal evolution of the rail pressure, mass flow rate, and energizing signal (a), and NOD time (b)

Fig. 5 Mass flow rate for different injection pressures and energizing times of Injector A

Fig. 6 Injected mass (a) and peak mass flow rate (b) for different energizing time and injection pressure of Injector A

Fig. 7 The mass flow rate for different injection pressure of Injector B

Fig. 8 Injection duration for different energizing time and injection pressure of Injector A and Injector B

Fig. 9 A schematic of the injection rate model

Fig. 10 The discharge coefficient against Reynolds number for injectors A and B

Fig. 11 Predicted injection rate profile and experimental data for Stage 1 (a) and Stage 2 (b) of injector B

Fig. 12 A schematic of the computational procedure with the injection model

Fig. 13 The injector needle valve open time for different energizing time and injection pressure of Injector B

Fig. 14 Injection rate profile for Injector B as a function of the time after the start of injection (ASOI); solid line:

experimental results; dashed line: model predictions

33
Fig. 15 Injection rate profile for Injector A as a function of the time after the start of injection (ASOI); solid line:

experimental results; dashed line: model predictions

Fig. 16 Injection rate profile for Injector C [50] as a function of the time after the start of injection (ASOI); solid line:

experimental results; dashed line: model predictions

Fig. 17 Injection rate profile for Injector D as a function of the time after the start of injection (ASOI); solid line:

experimental results; dashed line: model predictions

Fig. 18 Computational grids at TDC

Fig. 19 Variation of simulated in-cylinder pressure and heat release rate with crank angles for the baseline case

(IMEP=6bar) with different mesh sizes under 1500 rpm conditions

Fig. 20 Predicted injection rate profiles for the Diesel engine under different operating conditions

Fig. 21 Comparison of the experimental and simulation results for the Diesel engine under 1500 rpm conditions

(Solid line: experiments, dashed line: numerical simulations)

Fig. 22 Comparison of the experimental and simulation results for the Diesel engine under 1200 rpm and 1800 rpm

conditions (Solid line: experiments, dashed line: numerical simulations)

Fig. 23 Distribution of equivalence ratio and spray parcels at different injection pressures at IMEP= 5 bar (vertical

dashed line indicating the liquid penetration length)

Fig. 24 In-cylinder temperature distributions (IMEP= 5 bar)

34
lamp

Electronic box

EFS 8109

Electrovalve

Scale

Fig. 1 Sketch of injection rate measurement system

35
(a) Injector (b) Mono Injection Qualifier

Fig. 2 The injection rate experimental system

36
4 90
Energizing Time=0.9ms 20 injections per cycle Energizing Time=0.9ms
ROI and Command Signal

Flow mass rate (mg/ms)


Injection Pressure 120MPa 2 75 Injection Pressure 120MPa
3
1 60

2 Energizing 45
0
1046 1048 1050
ROI
30
1
15

0 0
0 500 1000 1500 2000 -0.3 0 0.3 0.6 0.9 1.2 1.5
Time (ms) ASOI (ms)

(a) (b)

Fig. 3 Temporal evolution of the instantaneous energizing current and ROI signals for 20 injections (a), and mean

averaged mass flow rate from the 20 injections (b) for Injector A.

37
160 120 0.5
Energizing Time=0.9ms Injection Pressure 120MPa Injector A 80MPa

Flow Mass Rate (mg/ms)

NOD (ms)
140 100 0.4 100MPa
Rail Pressure (MPa)

Nozzle Open Delay


120MPa
120 80 0.3

NOD (ms)NOD (ms)


140MPa

100 60 0.2
Nozzle Close Delay Injector B 50MPa
80 Rail Pressure 40 70MPa
Flow Mass Rate 0.4
90MPa
60 Energizing Signal 20 110MPa
Injection Duration 0.3 130MPa
40 0 150MPa
0 1 2 3 4 5 6 0.2
Time (ms) 0 0.4 0.8 1.2 1.6 2
Energizing Time (ms)

(a) (b)

Fig. 4 Temporal evolution of the rail pressure, mass flow rate, and energizing signal (a), and NOD time (b)

38
80 80
Injection Pressure=80MPa Injector A Injection Pressure=100MPa Injector A
Mass Flow Rate (mg/ms)

Mass Flow Rate (mg/ms)


0.5ms 0.5ms
60 0.7ms 60 0.7ms
0.8ms 0.8ms
0.9ms 0.9ms
40 1.1ms 40 1.1ms
1.3ms 1.3ms
1.5ms 1.5ms
1.8ms 1.8ms
20 20
Fluctuation Fluctuation

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Time (ms) Time (ms)

(a) (b)

80 80
Injection Pressure=120MPa Injector A Injection Pressure=140MPa Injector A
Mass Flow Rate (mg/ms)

Mass Flow Rate (mg/ms)


0.5ms 0.5ms
60 0.7ms 60 0.7ms
0.8ms 0.8ms
0.9ms 0.9ms
40 1.1ms 40 1.1ms
1.3ms 1.3ms
1.5ms 1.5ms
1.8ms 1.8ms
20 20
Fluctuation Fluctuation

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Time (ms) Time (ms)

(c) (d)

Fig. 5 Mass flow rate for different injection pressures and energizing times of Injector A

39
160 80
80MPa Injector A Injector A
100MPa

Max Flow Rate (mg/ms)


Injected Mass (mg/cyc)

120MPa
120 140MPa 60

80 40

80MPa
100MPa
40 20 120MPa
140MPa

0 0
0.4 0.8 1.2 1.6 2 0.4 0.8 1.2 1.6 2
Energizing time (ms) Energizing time (ms)

(a) (b)

Fig. 6 Injected mass (a) and peak mass flow rate (b) for different energizing time

and injection pressure of Injector A

40
60
Injector B Energizing Time=1.4ms

Mass Flow Rate (mg/ms)


50 Injection pressure=150MPa

40

30

20
Injection pressure=50MPa
10
1 2 3 4 5
0
0 0.4 0.8 1.2 1.6
ASOI (ms)

Fig. 7 The mass flow rate for different injection pressure of Injector B

41
2.4 2
injector A injector B
2
Injection duration (ms)

Injection duration (ms)


1.6

1.6 t id=ET t id=ET


1.2
1.2
50MPa
0.8
0.8 70MPa
80MPa 90MPa
100MPa 0.4 110MPa
0.4 120MPa 130MPa
140MPa 150MPa
0 0
0 0.4 0.8 1.2 1.6 0 0.4 0.8 1.2 1.6
Energizing Time (ms) Energizing Time (ms)

(a) (b)

Fig. 8 Injection duration for different energizing time and injection pressure of Injector A and Injector B

42
ROI

t5
t1 t2 t4 TIME

topen tclose

Fig. 9 A schematic of the injection rate model

43
0.9

Discharge Coefficient Cd
Cd ofCd of Injecor
Injector A A
0.85 Cd ofCd
Injector B B
of Injecor

0.8

y = -0.3128 x2+0.1867 x+0.8057


0.75

0.7
0.3 0.4 0.5 0.6 0.7 0.8
10000/Re

Fig. 10 The discharge coefficient against Reynolds number for injectors A and B

44
4 7.0
50MPa Injector B Solid Line: Exp Dashed Line: Fiting
Mass Flow Rate (mg/ms)

Mass Flow Rate (mg/ms)


70MPa
6.0 C2=4.13
90MPa
3 110MPa C2=3.71
130MPa 5.0 150MPa C2=3.22
150MPa C2=2.67
2 4.0
C2=2.07

3.0 C=1.30
C2=1.30
Solid line : Exp
1
Dashed line: Fiting
2.0 50MPa y = -0.8111 x2+2.3666 x+C2
3 2
Injector B y = a1 x +b1 x +c1x Single hole
0 1.0
0 0.1 0.2 0.3 0.4 0.2 0.4 0.6 0.8 1 1.2 1.4
ASOI (ms) ASOI (ms)

(a) (b)

Fig. 11 Predicted injection rate profile and experimental data for Stage 1 (a) and Stage 2 (b) of injector B

45
Input ρ f m fi  P D n ν f

uth Cd Gmax

a1 b1 c1 a2 b2 c2
topen
a4 a5 c5 tid

= topen + ∆t
tid
m f = ∫ G ( t )dt
0

∆ =m f i − m f
topen

∆ ≤ error

Output t id & Velocity profile

Fig. 12 A schematic of the computational procedure with the injection model

46
2
injector B
t open=ET
1.6

t open (ms)
1.2

0.8 50MPa
70MPa
90MPa
0.4 110MPa
130MPa
150MPa
0
0 0.4 0.8 1.2 1.6
Energizing Time (ms)

Fig. 13 The injector needle valve open time for different energizing time and injection pressure of Injector B

47
60 60
Injection Pressure=50MPa Injection Pressure=70MPa
Mass Flow Rate(mg/ms)

Mass Flow Rate(mg/ms)


50 50

40 40

30 30

20 20

10 10
0.20.3 0.4 0.5 0.6 0.8 1.0 1.2 1.4ms 0.2 0.3 0.4 0.5 0.6 0.8 1.0 1.2 1.4ms
0 0
0 0.3 0.6 0.9 1.2 1.5 1.8 0 0.3 0.6 0.9 1.2 1.5 1.8
ASOI (ms) ASOI (ms)

(a) (b)

60 60
Injection Pressure=90MPa Injection Pressure=110MPa
Mass Flow Rate(mg/ms)

50 Mass Flow Rate(mg/ms) 50

40 40

30 30

20 20

10 10
0.2 0.3 0.4 0.5 0.6 0.8 1.0 1.2 1.4ms 0.2 0.3 0.4 0.5 0.6 0.8 1.0 1.2 1.4ms
0 0
0 0.3 0.6 0.9 1.2 1.5 1.8 0 0.3 0.6 0.9 1.2 1.5 1.8
ASOI (ms) ASOI (ms)

(c) (d)

60 60
Injection Pressure=130MPa Injection Pressure=150MPa
Mass Flow Rate(mg/ms)

Mass Flow Rate(mg/ms)

50 50

40 40

30 30

20 20

10 10
0.2 0.3 0.4 0.5 0.6 0.8 1.0 1.2 1.4ms 0.20.3 0.4 0.5 0.6 0.8 1.0 1.2 1.4ms
0 0
0 0.3 0.6 0.9 1.2 1.5 1.8 0 0.3 0.6 0.9 1.2 1.5 1.8
ASOI (ms) ASOI (ms)

(e) (f)

Fig. 14 Injection rate profile for Injector B as a function of the time after the start of injection (ASOI); solid line:

experimental results; dashed line: model predictions

48
100 100
Injection Pressure=80MPa Injection Pressure=100MPa
Mass Flow Rate (mg/ms)

Mass Flow Rate (mg/ms)


Injector A Injector A
80 Solind Line: Exp 80 Solind Line: Exp
Dashed Line: Sim Dashed Line: Sim

60 60

40 40

20 20
0.5 0.7 0.9 1.1 1.3 1.5 1.8ms 0.5 0.7 0.9 1.1 1.3 1.5 1.8ms

0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5


ASOI (ms) ASOI (ms)

(a) (b)

100 100
Injection Pressure=120MPa Injection Pressure=140MPa
Mass Flow Rate (mg/ms)

Mass Flow Rate (mg/ms)


Injector A Injector A
80 Solind Line: Exp 80 Solind Line: Exp
Dashed Line: Sim Dashed Line: Sim

60 60

40 40

20 20
0.5 0.7 0.9 1.1 1.3 1.5 1.8ms 0.5 0.7 0.9 1.1 1.3 1.5 1.8ms

0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5


ASOI (ms) ASOI (ms)

(c) (d)

Fig. 15 Injection rate profile for Injector A as a function of the time after the start of injection (ASOI); solid line:

experimental results; dashed line: model predictions

49
100 100
Injector C Fuel : Diesel Injector C Fuel : RME
Mass Flow Rate (mg/ms)

Mass Flow Rate (mg/ms)


80 80

60 60

40 40
80MPa 80MPa
20 100MPa 20 100MPa
120MPa 120MPa
140MPa 140MPa
0 0
0 1 2 3 4 0 1 2 3 4
ASOI (ms) ASOI (ms)

(a) (b)

Fig. 16 Injection rate profile for Injector C [50] as a function of the time after the start of injection (ASOI); solid

line: experimental results; dashed line: model predictions

50
30
50MPa 75MPa 100MPa

Mass Flow Rate (mg/ms)


25 Injector D

20

15

10

0
0 0.4 0.8 1.2 0 0.4 0.8 1.2 0 0.4 0.8 1.2
ASOI (ms)

Fig. 17 Injection rate profile for Injector D as a function of the time after the start of injection (ASOI); solid line:

experimental results; dashed line: model predictions

51
Coarse mesh Fine mesh

Fig. 18 Computational grids at TDC

52
12 1200
Coarsh mesh

Heat Release Rate (J/°CA)


In-cylinder Pressure (MPa)
10 Fine mesh 1000

8 800

6 600

4 400

2 200

0 0
-30 -20 -10 0 10 20 30
Crank Angle (°CA ATDC)

Fig. 19 Variation of simulated in-cylinder pressure and heat release rate with crank angles for the baseline
case (IMEP=6bar) with different mesh sizes under 1500 rpm conditions

53
Injection Pressure =100MPa
80 80
mfuel=31.18mg/cyc 120MPa
Fuel mass=38.9 mg/cycle
Mass Flow Rate (mg/ms)

Mass Flow Rate (mg/ms)


mfuel=38.90mg/cyc 100MPa
mfuel=45.88mg/cyc 80MPa
60 mfuel=52.10mg/cyc 60

40 40

20 20

0 0
0 0.3 0.6 0.9 1.2 0 0.3 0.6 0.9 1.2
ASOI (ms) ASOI (ms)

(a) (b)

Fig. 20 Predicted injection rate profiles for the Diesel engine under different operating conditions

54
12 1200 500 1000
IMEP=4.0bar Rated speed=1500 r/min

Heat Release Rate (J/°CA)


In-cylinder Pressure (MPa)

Rated speed=1500 r/min IMEP=5.0bar


10 Injection pressure=80MPa 1000 Injection pressure=80MPa
IMEP=6.0bar 400 800
IMEP=7.0bar
8 800

NOx (ppm)

HC (ppm)
300 600
6 600
200 400
4 400

100 200
2 200

0 0 0 0
-30 -20 -10 0 10 20 30 4 5 6 7
Crank Angle (°CA ATDC) IMEP (bar)

(a) (b)

12 1200 500 1000


IMEP=4.0bar Heat Release Rate (J/°CA)
Rated speed=1500 r/min
In-cylinder Pressure (MPa)

Rated speed=1500 r/min IMEP=5.0bar


10 1000 Injection pressure=100MPa
Injection pressure=100MPa IMEP=6.0bar 400 800
IMEP=7.0bar
8 800
NOx (ppm)
HC

HC (ppm)
300 600
NOx
6 600
200 400
4 400

100 200
2 200

0 0 0 0
-30 -20 -10 0 10 20 30 4 5 6 7
Crank Angle (°CA ATDC) IMEP (bar)

(c) (d)

12 1200 500 1000


IMEP=4.0bar Rated speed=1500 r/min
Heat Release Rate (J/°CA)
In-cylinder Pressure (MPa)

Rated speed=1500 r/min IMEP=5.0bar


10 1000 Injection pressure=120MPa
Injection pressure=120MPa IMEP=6.0bar 400 800
IMEP=7.0bar
8 800
NOx (ppm)

HC
HC (ppm)

300 600
NOx
6 600
200 400
4 400

100 200
2 200

0 0 0 0
-30 -20 -10 0 10 20 30 4 5 6 7
Crank Angle (°CA ATDC) IMEP (bar)

(e) (f)

Fig. 21 Comparison of the experimental and simulation results for the Diesel engine under 1500 rpm conditions

(Solid line: experiments, dashed line: numerical simulations)

55
12 1200 1000 500
IMEP=4.4bar Rated speed=1200 r/min HC

Heat Release Rate (J/°CA)


In-cylinder Pressure (MPa)
Rated speed=1200 r/min
10 IMEP=6.0bar 1000 Injection pressure=120MPa NOx
Injection pressure=120MPa 800 400
IMEP=7.7bar

8 800

NOx (ppm)

HC (ppm)
600 300
6 600
400 200
4 400

200 100
2 200

0 0 0 0
-30 -20 -10 0 10 20 30 4 5 6 7 8
Crank Angle (°CA ATDC) IMEP (bar)

(a) (b)

12 1200 600 1200


IMEP=4.0bar Rated speed=1800 r/min HC
Heat Release Rate (J/°CA)
In-cylinder Pressure (MPa)

Rated speed=1800 r/min


10 IMEP=6.2bar 1000 500 Injection pressure=120MPa NOx 1000
Injection pressure=120MPa
IMEP=8.0bar
8 800 400 800
NOx (ppm)

HC (ppm)
6 600 300 600

4 400 200 400

2 200 100 200

0 0 0 0
-30 -20 -10 0 10 20 30 4 5 6 7 8
Crank Angle (°CA ATDC) IMEP (bar)

(c) (d)

Fig. 22 Comparison of the experimental and simulation results for the Diesel engine under 1200 rpm and 1800 rpm

conditions (Solid line: experiments, dashed line: numerical simulations)

56
80MPa
100MPa
120MPa

-2 °C ATDC 0 °C ATDC 2 °C ATDC 5 °C ATDC

Fig. 23 Distribution of equivalence ratio and spray parcels at different injection pressures at IMEP= 5 bar (vertical

dashed line indicating the liquid penetration length)

57
80MPa
100MPa
120MPa

5 °C ATDC 6 °C ATDC 8 °C ATDC 10 °C ATDC

Fig. 24 In-cylinder temperature distributions (IMEP= 5 bar)

58
Table 1 Specifications of the major experimental apparatus
Apparatus Model Specification
Injection system Common rail injection P<200 MPa, Rail volume: 30 cm3
Current Sensor Tektronix A622 Current <100 A Voltage: 10~100 mV
Mono injection qualifier EFS8246 0-600 mm3
Data acquisition Yokagawa DL750 Sample frequency: 50,000 Hz
Pressure Sensor Kistler 4067C3000 0-300 MPa
Amplifier Kistler 4618 -

59
Table 2 Nozzle parameters used in the experiments
Injector Injector A Injector B [36]
No. of Nozzle Holes 7 8
Hole Diameter (mm) 0.176 0.133
Spray Angle (o) 155.0 120.0
Injection Pressure (MPa) 80/100/120/140 50/70/90/110/130/150
Energizing Time (ms) 0.5~1.8 0.2~1.4

60
Table 3 Summary of key parameters for injector B
Injection Diameter Maximal Flow Velocity
Re Cd
Pressure(MPa) (mm) Rate(mg/ ms) m/s
50 0.133 2.95 329.2240 13258.97 0.768
70 0.133 3.56 394.3522 15881.91 0.800
90 0.133 4.08 450.1544 18129.26 0.813
110 0.133 4.54 499.7643 20127.21 0.821
130 0.133 4.96 544.8759 21944.01 0.834
150 0.133 5.35 586.5280 23621.48 0.829

61
Table 4 Physical properties of the fuels studied in the Injector C case [50]
Fuel property (20 oC ) Diesel RME
Density ( g/cm )
3
0.8234 0.8825
Kinematic viscosity (cSt) 4.59 6.57

62
Table 5 Engine specification and setup for numerical simulation

Engine parameters Value Simulation parameters Value


Displaced (L) 1.325 Engine speed (rpm) 1200/1500/1800
Engine Speed 1200~1800 IVC timing ( ATDC)
o
-145.5
No. of cylinders 4 EVO timing (o ATDC) 112.5
Bore (cm) 11.4 Swirl ratio 1.9
Stroke (cm) 13.0 Initial in-cylinder temperature 327
Connecting 21.6 Initial in-cylinder pressure (MPa) 0.103
Compression ratio 18.0:1 NOD time (ms) 0.32
Aspiration type Naturally Initial fuel temperature (K) 320.15
Piston bowl Open-crater-type Spray cone angle (o) 155.0
Rated power 147KW@1600r/min Injector hole diameter (mm) 0.176
Fuel supply system High-pressure Injection timing (o ATDC) -7
Injector BOSCH Fuel mass (mg/cycle) 31.18~52.10
Fuel Diesel Injection pressure (MPa) 80/100/120

63
Table 6 Injection duration and peak mass flow rate under different conditions

Load Fuel mass Injection Pressure Injection Duration Peak flow rate
(bar) mg/cyc (MPa) (ms) (mg/ ms)
80 0.924 45.67
4.0 31.18 100 0.832 50.93
120 0.764 55.78
80 1.092 46.86
5.0 38.90 100 0.985 52.19
120 0.905 56.77
80 1.239 47.80
6.0 45.88 100 1.119 52.98
120 1.029 57.82
80 1.366 48.31
7.0 52.10 100 1.235 53.76
120 1.137 58.41

64

View publication stats

You might also like