You are on page 1of 19

Energy 116 (2016) 470e488

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

CFD analysis on effect of localized in-cylinder temperature on nitric


oxide (NO) emission in a compression ignition engine under
hydrogen-diesel dual-fuel mode
Venkateswarlu Chintala, K.A. Subramanian*
Engines and Unconventional Fuels Laboratory, Centre for Energy Studies, Indian Institute of Technology-Delhi, New Delhi, 110 016, India

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 12 March 2016
Received in revised form
25 September 2016
Accepted 27 September 2016

Utilization of hydrogen in a compression ignition (CI) engine (7.4 kW rated power) under dual-fuel mode
could reduce all carbon based emissions however it emits high NOx (oxides of nitrogen) emission due to
high localized in-cylinder temperature during combustion. The present study is aimed at analysis of
effect of localized (burned zone) in-cylinder temperature on formation of NO emission using theoretical
(two zone model) and Computational Fluid Dynamics (CFD) simulations. Localized in-cylinder peak
temperature (in burned zone) increased from 2278.2 K with base diesel mode to 2402.7 K with dual-fuel
mode (16.7% hydrogen energy share). Nitric oxide (NO) emission formed mainly during premixed
combustion phase about 363 to 376 crank angle. The NO emission at 16.7% hydrogen energy share with
experimental test, two zone model, and CFD simulation are 914 ppm, 1208 ppm, and 1382 ppm. The
simulation results are inline with the experimental results with the error band of 15%e23%. It is well
established through this study that formation of NO emission at source level is strong function of
localized in-cylinder temperature and its distribution pattern in the combustion chamber.
2016 Elsevier Ltd. All rights reserved.

Keywords:
Computational uid dynamics (CFD)
Dual-fuel engine
Hydrogen
Oxides of nitrogen (NOx) emission

1. Introduction
Compression ignition (CI) engines are widely used in transportation sector, agriculture sector, and electrical power generation
sector. India is a diesel driven economy as diesel consumption is
almost six times higher than gasoline consumption [1]. In major
urban cities, the diesel fuelled automotive vehicles emit high level
of particulate matter which is higher than the permissible level.
Fuel economy improvement along with emissions reduction is a
driving force for enhancement of sustainability of these sectors.
Hydrogen utilization in diesel vehicles could be a solution to these
problems. In addition to this, the use of hydrogen as an alternative
fuel in CI engines would also satisfy the energy demand due to its
abundant availability in the universe [2,3]. Hydrogen could be
produced from water or any hydrocarbon feed stock through
different strategies [2]. Hydrogen addition in a CI engine under
dual-fuel mode enhances its thermal efciency and reduces carbon
based emissions including CO (carbon monoxide), HC (hydrocarbon), carbon dioxide (CO2), methane (CH4) and PM (particulate

* Corresponding author.
E-mail address: subra@ces.iitd.ernet.in (K.A. Subramanian).
http://dx.doi.org/10.1016/j.energy.2016.09.133
0360-5442/ 2016 Elsevier Ltd. All rights reserved.

matter) [4e6]. However, the dual-fuel engines face a major problem of high level of NOx (oxides of nitrogen: NO NO2) emission.
Formation of NOx emission could be propagated mainly by high
temperatures of fuel-air charge, excess oxygen concentration and
oxidation reaction time. Therefore, NOx emission is the highest in
zones close to stoichiometric and slightly air-rich conditions. The
hydrogen fuelled dual fuel engines operate with high adiabatic
ame temperature which increases the NOx emission drastically.
It is reported in literature that NOx emission increases with increase in in-cylinder temperature and the peak NOx level occur at
combustion peak temperatures which occurs between start of
combustion and in-cylinder peak pressure [7]. It is also reported
that almost all NOx formation occurs within 20 CA (crank angle)
from start of combustion [7]. In case of hydrogen based dual-fuel CI
engines, with increasing hydrogen energy share, NOx emission increases signicantly at high and medium loads while the emission
decreases at low loads [5,6,8]. Experimental investigations carried
out by Dhole et al. reported that NOx emission increased from 1.8 g/
kWh with base diesel mode to 0.67 g/kWh with 50% hydrogen
volume substitution (under dual-fuel mode) at 13% load in a CI
engine (62.5 kW rated power at 1500 rpm) [9]. This could be due to
increase in fraction of water content (H2O) which decreases the in-

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

cylinder peak temperature with hydrogen addition [10]. However,


at high load when more than 1% of hydrogen was substituted, NOx
emission increases as a result of faster combustion with hydrogen
due to high diffusivity and high ame speed [11]. It is generally
reported in literature that combustion with hydrogen proceeds
under wider range of air fuel ratios. If combustion proceeds with
rich air-hydrogen mixture, it results in increase in in-cylinder
temperature and NOx emission formation [12]. NOx emission in
hydrogen fuelled dual fuel engine is higher at lower speed than
high speed [13]. At high engine speed, NOx emission is low due to
less residence time. The summary of literature information on
variation of NOx emission from hydrogen dual-fuel engines is given
in Table 1.
It is emerged from the above literature review that NOx emission
is a major pollutant from hydrogen dual-fuel engines. Various NOx
emission reduction technologies or retrot devices at engine tail
pipe including Lean NOx Trap (LNT) and Selective Catalytic Reduction (SCR) are available. However, inadequate information is reported on the emission reduction technologies at sources level such
as water injection, injection timing retardation, exhaust gas recirculation (EGR) in case of hydrogen dual-fuel engines. Hence, a
proper insight on NOx emission formation at source level (inside
the combustion chamber of dual-fuel engine) would provide more
opportunities for the emission reduction at source level. In order to
address this issue, number of experimental research work needs to
be carried out on the hydrogen dual-fuel engines. But the experimental investigation with change of various input parameters for
the NOx emission reduction would be time consuming and it is also
difcult to conduct experimental tests under wider operating
conditions. The measurement of in-cylinder temperature at spatial
distribution in the combustion chamber is complex. NOx emission
with respect to in-cylinder temperature can be analysed using the
computational uid dynamics (CFD) and this would reduce number
of experimental tests to be conducted for understanding the NOx
emission formation at source level and also lowering the development cost and time signicantly. In these simulation models, the
effect of various design parameters (combustion chamber geometry, piston bowl geometry, compression ratio and fuel injection
timing etc.) and operating parameters (speed, intake air ow rate
and fuel ow rate etc.) on the emission formation could be estimated in fast and non-expensive way [20,21]. Sahin and Durgun
developed an emission model based on the chemical equilibrium

471

and kinetics of NO to determine the pollutant concentrations


within each zone and the whole of the cylinder [20]. Saravanan
et al. developed a correlation to predict thermal NOx emission
formation in a CI engine based with input of physico-chemical
properties (cetane number, density) and engine design parameters (bore, stroke, compression ratio, fuel injection timing) [22].
NOx emission formation in CI engines is strongly depends on high
in-cylinder temperature during premixed and diffusion combustion [23]. The in-cylinder temperature can be reduced using control
strategies for NOx emission reduction but on the other hand it is not
desirable as it may lead to reduction in thermal efciency of the
engine [6]. Hence, it is a challenge to reduce NOx emission at source
level without any other adverse effects on performance of the
engine.
Several studies were carried out on prediction of NO emission in
conventional CI engines by employing Zeldovich mechanism in
different zones of combustion chamber [24e26]. Instantaneous NO
emission concentration could be determined using in-cylinder
temperature and species concentration for each zone [27,28]. NO
emission in a hydrogen dual-fuel CI engine is generally predicted
using the extended Zeldovich mechanism based on the global incylinder temperature [8]. Modelling of NOx emission in a conventional diesel engine is complex due to combustion with heterogeneous charge [27]. However, it is still a challenging task for
modelling of dual-fuel CI engine where combustion proceeds with
both liquid diesel and gaseous fuels (hydrogen/compressed natural
gas/biogas/syngas/liqueed natural gas). Very few studies on
theoretical or CFD simulation in dual-fuel engines are available in
the literature. The few studies are mostly focused on natural gas
based dual-fuel engines but the studies on hydrogen based dualfuel engines are scanty. For example, Mattarelli et al. predicted
NO emission in a natural gas based dual-fuel engine (turbocharged
direct injection diesel engine) using KIVA code [29]. Similarly,
Maghbouli et al. used KIVA code for NOx emission prediction in a
natural gas based dual-fuel engine (multi cylinder turbo charged
diesel engine) [26]. Miao and Milton conducted both experimental
and computational study on a natural gas dual-fuel engine to understand the complications in dual-fuel combustion (i.e., ame
development, the relative contribution of the individual-fuels, and
the emission formation) [30]. A three-zone heat-release rate model
for natural gas dual-fuel engine was developed on the basis of
previously derived three-zone models for diesel engines [31]. A

Table 1
Literature details of NOx emission variation in hydrogen dual-fuel engines.
Reference

Engine details

Effect on NOx emission

Singh Yadav et al. [14]

Nc 1, CR 16.5:1,
BMEP 5.3 bar
Nc 4, CR 17:1,
BMEP 7.9 bar
Nc 4, CR 18.2:1,
BMEP 9.2 bar
Nc 6, CR 16.5:1,
BMEP 14.6 bar

 Increased from 440 ppm without hydrogen to 470 ppm with 0.04 kg/h hydrogen at 80% load
and 1500 rpm
 Increased from 2200 ppm without hydrogen to 2900 ppm with 7.5% hydrogen at 2200 rpm
 Decreased from 1370 ppm without hydrogen to 1010 ppm with 7.5% hydrogen at 1000 rpm
 Increased about 29% with 25.3% hydrogen at 5 bar BMEP and 2500 rpm
 Insignicant effect at low load and speed
 Increased from 4.8 g/kWh with 2% volume hydrogen to 6.4 g/kWh with 6% volume at 70% load
and 1200 rpm
 Decreased from 7 g/kWh with 2% volume hydrogen to 0.8 g/kWh with 6% volume at 10% load
and 1200 rpm
 At low load, the addition of hydrogen had mild effect on the emission
 At medium to high loads, the addition of a small amount of hydrogen (<3e5%) slightly reduced
the emission
 At high load, signicantly increased with the addition of 4% volume of hydrogen
Increased from 810 ppm without hydrogen to 1211 ppm with 0.15 kg/h hydrogen ow rate at 80% load

se and Ciniviz [15]


Ko
Christodoulou and
Megaritis [13]
Liew et al. [16]

Liew et al. [17]

Nc 6, CR 16:1,
BMEP 16.3 bar

Banerjee and Bose [18]

Nc 1, CR 17.5:1,
BMEP 6.2 bar
Nc 6, CR 16.5:1,
BMEP 14.6 bar

Liu et al. [19]

 Increased from 4.57 g/kWh with 2% hydrogen volume hydrogen to 5.28 g/kWh with 4% hydrogen volume
at 70% load
 Decreased from 5.6 g/kWh with 2% hydrogen volume hydrogen to 1.8 g/kWh with 6% hydrogen volume
at 15% load

472

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

chemical kinetics combustion model with a partially stirred combustion model (KIVA3V2 code) was used to predict the knock and
performance characteristics of a natural gas dual-fuel engine [32]. A
CFD analysis was carried out to study the effect of biogas addition
on NOx emission formation for ve compression ratios [33]. CFD
analysis using chemical kinetic mechanism was performed to
evaluate the combustion with syngas in a supercharged dual-fuel
engine for various syngas compositions under lean conditions
[34]. Rao and Honnery predicted NOx emission in a direct injection
CI engine using multi-zone quasi-dimensional phenomenological
model [35]. The principle of conservation of energy could also be
employed for determining localized in-cylinder temperatures in
different zones of the combustion chamber. In this methodology,
complex and tedious iterative calculations could be avoided for
determining the temperatures. Knop et al. adapted extended
coherent ame model for predicting the NO emission in hydrogen
fuelled IC engines [36]. Masood et al. carried out their research
work on combustion and emission analysis of hydrogen-diesel
blends in a CI engine under dual-fuel mode with port and manifold injection systems [37,38]. CFD analysis of dual-fuel combustion
and emissions were carried out using the CFD software FLUENT
[37]. It is explored from the literature study that very few studies
were reported on theoretical/CFD simulations of dual-fuel combustion and emissions formation in methane based dual-fuel CI
engines. However, the information on NO emission prediction in CI
engines under hydrogen based dual-fuel mode is scanty in the
literature. As the combustion characteristics of hydrogen are
entirely different than methane, there is a need to study the combustion process and emission formation in hydrogen based dualfuel CI engines. The turbulence of gas motion inside the combustion chamber of a dual-fuel engine plays a vital role in determining
its performance characteristics. A better understanding of gas
motion, in-cylinder pressure distribution and localized in-cylinder
temperature inside the combustion chamber of hydrogen dual-fuel
engines using CFD simulation will be helpful in optimizing the
engine design parameters. Furthermore, in most of the previous
studies, NO emission was predicted based on the global/average incylinder temperature during combustion but not on the localized
in-cylinder temperature. As the temperature is local phenomenon,
one needs to determine the emission formation based on localized
phenomenon. Hence, in the present study, an attempt has been
made to predict the NO emission using localized (burned zone) incylinder temperature in a CI engine under hydrogen dual-fuel
mode. Experimental tests were also carried out on a single

cylinder direct injection CI engine to validate the simulation results


of NO emission.
2. Experimental test setup
Experimental test setup mainly consisted of (i) a CI engine with
loading system, (ii) intake air and fuels supply and measuring
systems, (iii) auxiliary gaseous fuel injection system, and (iv)
combustion diagnosis system. The schematic diagram of experimental test setup is shown in Fig. 1.
2.1. Test engine integrated with loading system and exhaust gas
emission analyzers
A single cylinder constant speed CI engine (7.4 kW rated power
at 1500 rpm) was modied to run under dual-fuel mode. Detailed
specications of the test engine are given in Table 2. The engine was
loaded with an eddy current dynamometer (maximum capacity:
80 kW). Exhaust gas was tapped from tail pipe of the engine for
analyzing the NO emission with Chemiluminescence Detector
(CLD) analyzer under different engine operating conditions.
2.2. Details of combustion analysis system
Combustion analysis system comprised of an inbuilt charge
amplier, voltage amplier, and data acquisition system with 14 bit
processor. The system was mainly attached with three kinds of
sensors; (i) piezoelectric pressure transducer (ii) optical encoder
and (iii) piezoelectric strain gauge. The piezoelectric pressure
transducer with nominal sensitivity 45 pC/bar was ush mounted
on the cylinder head of the engine, for capturing instantaneous incylinder pressure data during the engine operation. The optical
encoder (720 pulses/revolution) was mounted on one end of
crankshaft of the engine, for degrees crank angle (CA) measurement with accuracy of 0.1 CA. The piezoelectric strain gauge
pressure transducer (pressure measurement range: 0e2000 bar)
was tted on high-pressure diesel pilot fuel line for measuring the
pilot fuel injection pressure. The in-cylinder pressure analog signal
was amplied by the charge amplier and then, the analog signal
was converted to a digital signal from the data acquisition system
for further processing of the acquired data. The in-cylinder pressure
signal, in-line pilot fuel pressure signal, and TDC (top dead centre)
encoder signals are processed together by the combustion analysis
system. Pressure data was indexed with the angle position of the

Fig. 1. Schematic diagram of hydrogen dual-fuel engine.

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488


Table 2
Technical specications of the CI engine.
Description

Values

Type of engine
Number of cylinders
Displacement volume, cc
Rated power output, kW
Rated speed, rpm
Bore  Stroke, mm
Compression ratio
Intake valve opening and closing,
degree CA
Exhaust valve opening and closing,
degree CA

Compression ignition
1
947.4
7.4
1500
102  116
19.5:1
43 before TDC & 67 after BDC
87 before BDC & 39 after TDC

crankshaft, with reference to the compression TDC. The post processing software was used for processing of pressure-crank angle
data.
2.3. Details of intake air and fuels (diesel and hydrogen) measuring
systems
Both liquid and gaseous fuels were used in the study. Bharat
stage-IV diesel was used as pilot fuel, and hydrogen was used as
main fuel in the test engine. The physiochemical properties of the
fuels are given in Table 3. The liquid pilot fuel was supplied to the
engine with the unmodied conventional diesel injection system.
The pilot fuel was directly injected into the combustion chamber at
the pressure range from 250 bar to 500 bar. Conventional high
pressure liquid injector was used for the study. The pilot fuel consumption was measured using a high accuracy digital electronic
weighing balance while the main gaseous fuel (hydrogen) consumption was measured with Coriolis based gas mass ow meter as
shown in Fig. 1. The main gaseous fuel was supplied to the engine
using a newly developed gas management system. The intake air
system of the engine consists of an air lter assembly, a surge tank,
an air ow meter, and an intake manifold. The air ow meter was
tted in upstream of the intake manifold of the engine to measure
airow rate. A surge tank was tted in upstream of the air ow
meter in order to avoid signal noise due to pulsating ow of intake
air.
2.4. Auxiliary injection system for hydrogen gas
A timed manifold injection system was developed for injecting
hydrogen gaseous fuel into the intake manifold. The gas injection
system mainly consisted of an Electronic Control Unit (ECU), solenoid gas injector, ashback arrestor/ame arrestor, gas mass ow
meter, gas pressure reducer, and gas fuel lines with pressure gauges
and control valves. The compressed hydrogen gas was typically
stored in the cylinder at 200 bar pressure and the fuel was injected
into the engine's intake manifold using an electronic injector at
2 bar pressure. The pressure reducer valve, which was located at the

Table 3
Properties of the fuels used in the experimental study [39e42].
Fuel characteristics

Diesel

Hydrogen

Molecular structure
Lower heating value, MJ/kg
Stoichiometric air fuel ratio
Auto ignition temperature, K
Laminar burning velocity, m/s
Cetane number
Density, kg/m3

C12H26
44.05
14.5
530
0.3
51
821.5

H2
120
34.2
858
2.65e3.25
e
0.083

473

neck of the hydrogen cylinder, was used for reducing the gas
pressure from 200 bar to 2 bar. Flame arrestor/ash back arrestors
were integrated in the hydrogen gas fuel line at different locations
to arrest the backring of the hydrogen gaseous fuel.

2.5. Details of experimental tests


Experimental tests were conducted on the test engine under
diesel-hydrogen dual-fuel mode at 100% load (brake mean effective
pressure: 6.1 bar) and constant speed of 1500 rpm. Diesel fuel (an
ignition source) was directly injected into the combustion chamber
at the end of compression stroke whereas hydrogen was injected
into the intake manifold during suction stroke. Hydrogen was
injected into the intake manifold after the outlet valve closed (43
CA after TDC) in order to avoid scavenging losses. Start of hydrogen
injection was maintained constant as 43 CA after TDC throughout
the experimentation whereas the end of hydrogen injection was
varied with respect to the engine loading. These injection timings
were optimized based on better performance and lower emissions
for diesel-hydrogen dual-fuel operation [43]. Measurement values
range and resolutions of all the analysers/equipment used in the
experimental investigation are given in Table 4. All experimental
measurements, irrespective of the type of instrument used, possess
some uncertainty. Uncertainties of various parameters are determined using Eq. (1) and their values are given in Table 4. Model
calculations for uncertainty of hydrogen energy share are given in
Appendix.

"
q

vq
Dx
vx1 1

2

vq
Dx
vx2 2

2

vq
Dxn
vxn

2 #12

(1)

3. Methodology for prediction localized in-cylinder


temperature and the corresponding NO emission using two
zone model
In case of hydrogen based dual-fuel engines, NOx emission is the
major pollutant as discussed in the literature. Typically NOx emission from IC engines consists of NO (nitric oxide) and NO2 (nitrogen
di-oxide), in which NO would be the major share. Hence, in the
present study, an attempt has been made to predict the NO emission from a hydrogen based dual-fuel engine using two zone model
(localized in-cylinder temperature/burned zone temperature). The
entire combustion chamber is divided into two zones; (i) unburned
zone where hydrogen-air mixture is available, and (ii) burned zone
where diesel-hydrogen-air mixture is available as shown in Fig. 2.
As the diesel fuel injector has 5 number of holes (5 sprays), the total
burned zone is equal to the sum of 5 burned zones. It may be noted
that after the completion of pilot diesel injection, the combustion
chamber consists of burned and unburned zones only for the rest of
combustion duration. The assumptions for two zone modelling are
given below.
The species in combustion chamber (diesel, air, and hydrogen)
are considered as ideal gases
The sum of the volumes of Zone 1 and Zone 2 is equal to 1/5th of
the total volume of the combustion chamber
The system is considered as closed and therefore mass loss due
to blow-by/crevices is ignored
In-cylinder pressure is uniform in all zones, but in-cylinder
temperature varies with respect to the zone

474

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

Table 4
Details of measurement range, resolution, and accuracy of instruments/sensors with uncertainty.
Instrument/sensor name

Measuring parameter

Measuring range

Resolution

Accuracy

Uncertainty (%)

CLD analyzer

NOx emission

0-10,000 ppm

1 ppm volume

2.692

Piezoelectric pressure
sensor
Optical encoder

In-cylinder pressure

0-250 bar

Sensitivity: 45 pC/bar

<1000 ppm volume: 5 ppm


1000 ppm volume: 5% of value
0.3% to 0.6% of value

Air ow meter
Gas mass ow meter
Calculated parameter

Degree crank angle


Intake air volumetric ow rate
Gaseous fuel ow rate
Hydrogen energy share

720 pulses per


revolution
0-330 m3/h
0-10 kg/h
e

0.1 CA
3

0.2 m /h
0.1 kg/h
e

0.846
(Peak pressure)
e

2% of ow
1% of ow
e

1.7
1.4
0.33

Since; dhubz Cpubz dTubz

(6)

dQ ubz mubz Cpubz dTubz  Vubz dp

(7)

The above Eq. (7) could be rewritten as in Eq. (8),

dTubz

dQ ubz Vubz dp
mubz Cpubz

(8)

Differential unburned zone temperature with respect to degree


crank angle is given in Eq. (9).



dTubz
1
dQ ubz
dp

Vubz
mubz Cpubz
dq
dq
dq

(9)

Where
ubz
Vubz dp
g dV
dQ ubz
dq
dq

dq
g1

Fig. 2. Schematic diagram of different zones in the combustion chamber.

The mixture, during the combustion phase is divided into two


zones: burnt gases zone, and unburned gases (after the end of
diesel spray)
Each zone is considered as open system through which mass
transfer takes place. Law of conservation of energy could be applied
to each independent zone in order to derive the localized combustion temperature as given in Eqs. (2)e(9).

(10)

3.2. Zone 2 (burned zone: diesel-hydrogen-air mixture)


Similarly, the differential burned zone temperature (Tbz: localized in-cylinder temperature) could be determined by applying
rst law of thermodynamics to zone 2 (burned zone) as given in
Eqs. (11)e(17),

dQ bz dUbz p dVbz  dmbz hbz

(11)

dUbz mbz dhbz hbz dmbz  p dVbz Vbz dp

(12)

3.1. Zone 1 (unburned zone: hydrogen-air mixture)


First law of thermodynamics applied to zone 1 (unburned zone)
is given in Eq. (2),

dQ ubz dUubz p dVubz  dmubz hubz

(2)

Above two equations could be rearranged as in Eq. (13),

dQ bz mbz dhbz hbz dmbz  p dVbz  Vbz dp p dVbz


 dmbz hbz

The internal energy of any thermodynamic system is given in Eq.

(13)

(3)

Uubz mubz hubz  p Vubz

(3)

Differentiating the above Eq. (3),

Since; dhbz Cpbz dTbz

(14)

dQ bz mbz Cpbz dTbz  Vbz dp

(15)

dUubz mubz dhubz hubz dmubz  p dVubz Vubz dp


(4)
Substituting Eq. (4) in to Eq. (2)

dQ ubz mubz dhubz hubz dmubz  p dVubz  Vubz dp


p dVubz  dmubz hubz

(5)

dTbz

dQ bz Vbz dp
mbz Cpbz



dTbz
1
dQ bz
dp

Vbz
mbz Cpbz
dq
dq
dq

(16)

(17)

The average in-cylinder temperature of burned and unburned


zones could be determined using a simple arithmetic mean Eq. (18).

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488



dTaverage 1 dTbz dTubz

2 dq
dq
dq

(18)

475

used as a pilot fuel in the engine, spray penetration and mixture


entrainment could be determined using Eqs. (31) and (32)[44].

The stoichiometric combustion equations for diesel (CaHb) and


hydrogen gas (Hg) fuels separately are described as given in Eqs.
(19) and (20).

 
mH2air C5 rH2air rdiesel 0:5 dnozzle Sprpenetration vinjection tan
2
(31)


y
y
Cx Hy x O2 3:773N2 xCO2 H2 O
4
2

y
3:773 x N2
4

Spray penetration; Sprpenetration






DP 0:5
530
3:07
dnozzle t0:5
TH2air
rH2air

g
g
g
Hg O2 3:773N2 H2 O 3:773N2
4
2
4

(19)

(20)

For dual-fuel operation the stoichiometric combustion equation


is given as in Eq. (21),





ay bg
O2 3:773N2
a Cx Hy bHg ax
4




ay bg
ay bg
H2 O 2ax
3:773N2
axCO2
2
2

(32)

Where

C4 Constant; 0:39 ;

C5 Constant; 0:28

Dp Pressure gradient between fuel injection and in-cylinder


pressures.
Instantaneous combustion chamber volume (V) is given in Eqs.
(33) and (34),
Vq Vbz q Vubz q

(33)

pVq mbz Rbz Tbz q mubz Rubz Tubz q

(34)

(21)
Where

md
md mg

(22)

mg
b
md mg

(23)

Cp;


ni R 
Ai; 1 Bi;2 T Ci;3 T2 Di;4 T3 Ei;5 T4
mi

; kJ=kg  K
(35)

Stoichiometric air fuel ratio could be determined using Eq. (24)

AFs

Specic heat at constant pressure (Cp) of species could be


determined using Eq. (35),[7].

h 

 i
a x 4y b 4g Mair

(24)

a12:011x 1:008y b1:008g

Total mass of species (m) contained in the combustion chamber


is given in Eqs. (25)e(27),

mq mbz q mubz q
mbz q lbz


A F s 1 m q
f

mubz q mq  mbz q

2 Dp

Vdiesel
nnozzle holes

(36)

N O2

/ k2f

NO N

(37)

dNO
k1f N2 O k2f NO2 
dt

(38)

dN
k1f N2 O  k2f NO2 
dt

(39)

0:5

rdiesel

(29)

Volume of each pilot fuel spray

NO N

(27)

(28)
Volume of total pilot fuel injected=cycle; Vdiesel
mdiesel



rdiesel N 120

/ k1f

(26)

Details of pilot diesel fuel spray i.e., injection velocity, volume of


diesel fuel spray, spray penetration and air-hydrogen mixture
entrainment are given in Eqs. (28)e(32).

Pilot diesel fuel injection velocity; vinjection C4

O N2

From the above elementary reactions, the products (NO and N)


formation rates by neglecting reverse reactions could be written as
given in Eqs. (38) and (39).

(25)

In the present study, Zeldovich mechanism (Eqs. (36) and (37))


was used for theoretical modelling of NO formation in the CI engine
[45].

(30)

In case of a hydrogen dual-fuel engine, premixed hydrogen-air


mixture entrains into the injected pilot fuel spray. As diesel fuel

With steady state approximation, d [N]/dt 0; then the above


equations could be rewritten as in Eqs. (40)e(42)[45].

!
dNO
k1f N2 O
k1f N2 O k2f
O2 
dt
k2f O2 

(40)

dNO
2k1f N2 O
dt

(41)

Kp Po
dNO
2k1f ,
dt
RT
Where

!0:5
N2 O2 0:5

(42)

476

k1f

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488


1:82 10 ^14 exp  38370=T

4.1. Details of different types of grids used for the CFD study

(43)

p
N2 
x
_RT N2

(44)

p
x
_RT O2

(45)



DGoT
kp Po exp
_
RT

(46)

O2 

4. CFD simulation for prediction of NO emission in the engine


under dual-fuel mode
A three dimensional hexahedral mesh for the combustion
chamber geometry (Fig. 3) was created using GAMBIT 2.4.6 software and then numerical simulation was carried out using ANSYS
FLUENT software. Dynamic mesh model was used for motion of
piston where the shape of the domain is continuously changing
with respect to time.

CFD analysis of combustion and emission formation in internal


combustion engines is a quite complex due to involvement of dynamic grid (mesh) in transient scale. The process of dynamic
meshing associates with generation/destruction of mesh elements
with respect to motion of the piston (degree crank angle). Accuracy
of the CFD results strongly depends on the total number of elements per unit area of computational working domain (i.e., mesh
density). Mesh density is an important metric used to control the
accuracy (element type and shape). Assuming no singularities are
present, a high-density mesh will produce results with high accuracy. As the computational power increases exponentially with the
increasing mesh density, it requires a lot of computational memory
and cost. In addition to this, high density domain may cause
divergence problem during simulation due to complex geometry of
combustion chamber (particularly piston bowl) of the CI engine. On
the other hand, less density domain gives less accurate results.
Hence, trade-off needs to be done for choosing the optimum mesh
density for the working domain.
In GAMBIT software, mesh density is represented with node
spacing on the mesh face form. When node spacing (or mesh
density) is specied, GAMBIT applies the specication to all the
edges associated with any specied faces that are not meshed. In

Fig. 3. (a) Solid geometry of cylinder and piston bowl (b) Photograph indicating piston with its bowl (cef) Different types of hexahedral mesh for sector of 72 .

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

477

Table 5
Details of mesh types and its quality measures for a sector domain of 72. .
Mesh description

Node spacing @ interval


size (mesh density)

Number of nodes

Number of cells

Number of quadrilateral
interior faces

Aspect ratio

Skewness

Mesh
Mesh
Mesh
Mesh

2.0
1.6
1.2
0.8

23691
39924
52548
93637

19572
34637
47270
87371

58232
101378
136701
249626

3.2
2.5
1.8
1.6

0.01
0.02
0.04
0.05

type
type
type
type

1
2
3
4

the present study, among the three different ways (Interval count;
Interval size; shortest edge (%)), interval size was varied to control
the mesh density of the working domain as given in Table 5. Node
spacing was decreased from 2 (default value in GAMBIT) to 0.8
insteps of 0.4 (Table 5). Total number of intervals (n) on the edge
could be determined using Eq. (47) in the software. If n is a noninteger, GAMBIT rounds to the nearest whole number to determine the number of intervals on the edge. It may be noted that
there is no such thumb rules for the size of the elements relative to
the geometry because there are too many different factors that
affect the results. What element size is the best for the given model/
domain depends on what results one needs to capture at what
locations in the model. For example, more focus need to do be given
on piston bowl locations where the most of the combustion takes
place inside the cylinder. In view of complicity and computational
cost associated with engine simulations, Maghbouli et al. also carried out the CFD simulation work with less number of cells about
2347 cells, 5752 cells, and 13408 cells to understand the combustion characteristics in a dual-fuel engine [26]. It is known that for
the same cell/element count, hexahedral meshes will give more
accurate solutions, especially if the grid lines are aligned with the
ow. Hence, in the present study hexahedral mesh elements were
selected for the computational working domain.

Edge length
Total number of intervals n
spacing interms of interval size
(47)

turbulence kinetic energy (k) and its dissipation rate (Epsilon: ).


Discretion settings of the model along with its constants are given
in Table 9.

v
v
v
rk
rkui
vt
vxi
vxj

"

m vk
m t
Gk Gb  r  Ym
sk vxj
(48)

Where.
v
vt k Local derivative of turbulent kinetic energy
v
vxi kui Rate of change of kinetic energy per unit mass

to turbulence, and viscous stress


Gb Generation of turbulent kinetic energy due to buoyancy
Gk Generation of turbulent kinetic energy due to mean velocity gradients
Ym Contribution of uctuating dilatation due to turbulence
and dissipation rate

v
v
v
r2
r2ui
vt
vxi
vxj

"

Various numerical models (dynamic meshing model, k-epsilon,


discrete phase, species transportation with reactions, and emission
formation models) used for CFD analysis of the dual-fuel CI engine
combustion and NO emission formation. The summary of these
numerical models and the solution methods used in the study are
given in Tables 6 and 7. Boundary conditions employed to different
zones in the working domain is given in Table 8.
4.2.1. k-Epsilon turbulence model for dual fuel engines
A standard two-equation k-Epsilon model was used to determine both turbulent length and time scale by solving two separate
transport equations [46]. The model was proposed by Launder and
Spalding based on transport equations (Eqs. (48) and (49)) for the

mt v2
2
C12 Gk
k
s2 vxj

C32 Gb  C22 r
Where.

4.2. Numerical models and boundary conditions used in the study

due to

convection (or advection)


"
 #
mt vk
v
m sk vxj Transport of kinetic energy in ow eld due
vxj

22
k

(49)

v
vt 2

Rate of change of dissipation per unit mass (Local


derivative)
v
vxi 2ui Rate of change of dissipation per unit mass due to
convection
"
 #
mt v2
v
m s2 vxj Transport of dissipation rate in ow eld
vxj
due to turbulence, and viscous stress

These above equations are solved using CFD software. Constants, C1, C2, C3, Cm, s, and sk are considered for calculations are
as follows:

C1 1:44; C2 1:92; C3 1:3; Cm 0:09; s 1:3; and sk


1:0

Table 6
Summary of models used in CFD simulation.
Description

Models used

Solver
Dynamic mesh
Spatial discretization
Turbulence
Turbulence chemistry interaction
(Combustion)
Diesel pilot fuel injection
(spray model)
NO emission

Pressure based type (Transient)


Layering (In-cylinder options)
Second order upwind
Standard k-Epsilon
Eddy dissipation concept (EDC)
Discrete phase mode (DPM); Solid cone
Thermal NO and Prompt NO

4.2.2. Dynamic mesh model for dual fuel engines


In CFD simulation of engine operation, as valves and the piston
move, the mesh created for the geometry has to move in order to
simulate valves and piston position with respect to degree crank
angle motion. Piston and piston bowl movement are decided by the
length of stroke, connecting rod length and crank angle radius. In
the present CFD study, only piston motion simulation was carried
out in order to assess the combustion and emission characteristics
of hydrogen based dual-fuel engine. Boundary conditions for

478

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

General scalar quantity


V Control volume
r Fluid density
u Flow velocity vector
ug Mesh velocity of the moving mesh
x Diffusion coefcient
S Source term of

Table 7
Details of solution methods used in CFD simulation.
Description

Parameter

Method/Model/Value

Pressure-velocity
coupling

Scheme
Skewness correction
Neighbor correction
Gradient
Pressure
Density
Momentum
Turbulent kinetic energy
Turbulent dissipation rate
All species including O2,
N2, H2 and C7H16
Energy
Autoignition

PISO scheme
0
1
Least squares cell based
Standard
Second order upwind
Second order upwind
Second order upwind
Second order upwind
Second order upwind

Spatial
discretization

Second order upwind


Second order upwind

Table 8
Boundary conditions for different zones.
Description

Parameter

Value/Method

Cylinder wall/Cylinder
head/Piston

Momentum

Wall motion
Shear condition
Roughness
height (mm)
Roughness constant
Thermal condition

Stationary wall
No slip
0

Thermal

0.5
Heat ux

Value

k-epsilon model
Near-wall treatment
Cmu (Model constant)
C1-Epsilon (Model constant)
C2-Epsilon (Model constant)
TKE Prandtl Number (Model constant)
TDR Prandtl Number (Model constant)
Energy Prandtl Number (Model constant)
Wall Prandtl Number (Model constant)
Turbulent Schmidt Number (Model constant)

Standard (2 equation)
Standard wall function
0.99
1.44
1.92
1
1.3
0.85
0.85
0.7

dynamic motion of piston are given in Table 10. The general


transport equation (Eq. (50)) which is applicable for all model
equations including energy, species, and turbulence etc. was used
in the CFD simulation [46].

rdV

rdV

r u  ug $dA

rVn1  rVn
Dt

(51)

Where, n and n1 denote the respective quantity at the current


and next time level. The (n1)th time level volume (Vn1) is
determined using Eq. (52),

dV
Dt
dt

(52)

Where, dV/dt is the volume time derivative of the control volume. In order to satisfy the mesh conservation law, the volume time
derivative of the control volume is computed using Eq. (53).

Z
ug $dA

nf
X

ugj $Aj

(53)

The dot product ugj :Aj on each control volume face is computed
using Eq. (54)

Description

dV

dt

Table 9
Discretion settings of viscous model used in the study.

d
dt

d
dt

Vn1 Vn

Zone type

Wall
roughness

By using a rst-order backward difference formula, the time


derivative term in Eq. (50) could be written as in Eq. (51)

xV$dA

S dV

(50)

ugj $Aj

Table 10
Discretion settings (boundary conditions) of In-cylinder motion
analysis.
Description

Value

Crank shaft speed (rpm)


Starting crank angle (degrees)
Crank period (degrees)
Crank angle step size (degrees)
Crank radius (mm)
Connecting rod length (mm)
Piston pin offset (mm)
Piston stroke cutoff (mm)
Minimum valve lift (mm)

1500
330
720
0.2
58
232
0
0
0

(54)

Where.
nf Number of faces on the control volume
Aj j face area vector.

dVj Volume swept out by the control volume face j over the
time step Dt.
4.2.3. Combustion model for dual fuel engines
Combustion in the engine was modelled using eddy dissipation concept (EDC) model. In this model it assumes that
combustion reaction occurs in small turbulent structures (ne
scales). The length fraction of the ne scales is modelled as
given in Eq. (55),

a Ca

Where.

dVj
Dt


w 0:25
k2

(55)

Ca Volume fraction constant 2.1377.


w Kinematic viscosity.
Species are assumed to react in the ne structures over a time
scale as given in (Eq. (56),

t Ct

 0:5
w

(56)

Where Ct Time scale constant 0.4082.


Combustion at the ne scales was assumed to occur as a constant pressure reactor, with initial conditions taken as the current
species and temperature in the cell. Reactions proceed over the
time scale t, governed by the Arrhenius rates of Equation, and are

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

integrated numerically using the ISAT algorithm [46]. The source


term in the conservation equation for the mean species i, is
modelled as given in Eq. (57),

ra2 
Yi;
Ri

t 1  a3

 Yi

v
rYi V$rwYi V$Ji Ri Si
vt

(58)

4.2.4. Auto ignition model for dual fuel engines


Hardenburg model [46] was used for auto-ignition of diesel fuel
in the present study. The transport equation for an ignition of
species, Yig could be represented as Eq. (59).





vrYig
mt
V$ rwYig V$
VYig rSig
vt
Sct

(59)

The ignition delay period was calculated using the Hardenburg


and Hase correlation as given in Eq. (60). The boundary conditions
for the model are given in Table 11. The activation energy of fuel-air
mixture was taken from the earlier investigation on hydrogen
based dual-fuel engine [23,47], [48].

tid

 
ep

 
C1 0:22 Sp
1
1
21:2


exp Ea
RT 17; 190
p  12:4
6N
(60)

4.2.5. NO emission model for dual fuel engines


In the present CFD simulation study, the mass transport equation (Eq. (61)) for the NO species is solved by considering convection, diffusion, production and consumption of NO and related
species.

vrYNO
V$rwYNO V$rDVYNO SNO
vt

(61)

After computing the in-cylinder temperature in the combustion


chamber, Zeldovich mechanism was employed to determine the NO
emission. Nitric oxide (NO) emission in ppm is computed using Eqs.
(62)e(64)[46],

NO ppm

NO mole fraction  106


1  H2 O mole fraction

NO mole fraction

Table 11
Discrete setting conditions of autoignition model.

(57)

Where Yi; r Fine scale species mass fraction after reacting over
the time. t
The EDC model can incorporate detailed chemical mechanisms
into turbulent reacting ows. Multiple simultaneous chemical reactions can be modelled, with reactions occurring in the bulk phase
(volumetric reactions) and/or on wall surfaces. The species transport equation could be written as (Eq. (58).

(62)

479

Description

Value

Model
Fuel species
Pre-exponential
Pressure exponent
Activation energy (J/mol)
Cetane number

Hardenburg and Hase


C7H16 (n-heptane)
1.2
1.5
20840
51

Mixture molecular weight P

mass fraction of species


i molecular weight of species

(64)

4.3. Convergence criteria


In CFD analysis, for an assessment on the numerical accuracy of
the computed results, a mesh sensitivity analysis needs to be carried out. The error that can be quantied by mesh renement is
known as the discretization error. Grid convergence index (GCI)
method yields discretization error bands for the different considered variables and thus provides a sophisticated quantication of
mesh dependency. However, in the present study the GCI methodology was not used for grid independence test. Grid/mesh independence test was conducted on the important combustion
performance indicators of in-cylinder pressure and in-cylinder
peak pressures. In CFD analysis of engine processes, Fixed Time
Stepping method (Crank Angle Step method) was used to run the
simulations. For cold ow simulation where there is no combustion, the value of the Crank Angle Step Size should be either 0.2 or
0.25 CA. For cases involving combustion, Crank Angle Step Size
should be smaller i.e., about 0.1 CA. In the present study, initially
the time step size of 0.2 CA was adopted and then it will be reduced
automatically to 0.1 CA when combustion occurs which will be set
up later (after completion of combustion process). As in the engine
analysis, only combustion and power stroke is of interest, the CFD
simulation was carried out only from 30 CA before TDC (top dead
centre) to 60 CA after TDC (i.e. for 90 CA duration). Hence, total
536 time steps were adopted to capture the combustion process in
each category of simulation. With reference to earlier research
works carried out on engine simulations, the maximum number of
iterations was considered to be 50 [46]. The convergence criteria for
different controlling parameters are given in Table 12.
5. Results and discussion
5.1. Nitric oxide (NO) emission prediction using two zone model
Fig. 4 shows the in-cylinder temperature variation for different
zones in the engine under dual-fuel mode (16.7% hydrogen energy
share) at 100% load. It may be noted that the global in-cylinder
temperature was determined based on the measured in-cylinder
pressure data (experimental data) with respect to degree CA using ideal gas Eq. (65) whereas the average in-cylinder temperature
was determined using Eq. (18) (average of burned and unburned
zones temperatures). It is clearly observed from the simulation
results that burned zone temperature (localized in-cylinder

NO mass fraction  mixture molecular weight


30

(63)

480

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488


Table 12
Details of convergence criteria for residuals.
Residuals

Absolute criteria

Continuity
x-velocity
y-velocity
z-velocity
Energy
k
epsilon
Ignition

0.1
0.001
0.001
0.001
1e06
0.001
0.001
0.001

temperature) is signicantly higher than the global/average incylinder temperature during the combustion. The reason for this
signicant difference is due to localized phenomenon i.e., the incylinder temperature varies with respect to spatial coordinates in
the combustion chamber. At 16.7% hydrogen energy share, the incylinder peak temperature in burned zone is about 2379 K
whereas the global peak temperature (which is calculated with the
input data of measured in-cylinder pressure) is about 1840 K as
seen in Fig. 4. The unburned zone temperature is substantially
lower than the global as well as burned zone temperatures, due to
no oxidation reaction takes place in the unburned zone.
dp
 dV
dTglobal
dq
dq

dq
mairfuel charge  R

(65)

Fig. 4. In-cylinder temperature variation at different zones in the combustion


chamber.

It is well reported that nitric oxide (NO) emission formation


during the combustion mainly depends on three parameters; (i)
combustion temperature, (ii) oxygen content in the fuel-air charge,
and (iii) reaction time in the engine [38,49e51]. Abdelaal et al.
investigated the effect of oxygen concentration on NO emission in a
hydrogen dual-fuel engine (6 kW rated power at 2000 rpm) and
concluded that the NO emission increased signicantly with increase in oxygen concentration inside the combustion chamber
[52]. They stated that the main reason for NO emission increment is
due to increase in the in-cylinder temperature with increasing
oxygen concentration. Hence, it could be concluded that thermal
NO route plays a vital role in shooting up of NO emission level in
case of hydrogen dual-fuel engines. Heywood supports the fact that
the NO emission formation is exponentially proportional to the incylinder temperature and directly proportional to the oxygen
concentration [7]. In case of hydrogen dual-fuel engines, the pilot
diesel (small quantity of liquid fuel) is injected at the end of the
compression stroke to ignite the hydrogen gaseous-fuel air mixture
[43]. Prior to the diesel fuel injection, the hydrogen fuel-air mixture
undergoes pre-ignition chemical reactions during compression
stroke, and subsequently result in formation of active radicals and
partial combustion products [52]. As soon as the liquid diesel is
injected into the combustion chamber, a distribution of equivalence
ratios develops across the pilot spray. Mainly, three regions would
form around the diesel fuel spray inside the combustion chamber.
These regions are (i) fuel rich region: the fuel spray core which is
relatively rich, (ii) nearly stoichiometric region: at downstream of
the core where combustion commences and (iii) lean region: the
fuel outside the spray boundary is mixed leaner than the lean
combustion limit [7]. Accordingly, NO formation in dual-fuel engine
follows both routes: the thermal mechanism and the prompt
mechanism [52]. Thermal NO is formed as a result of the intense
combustion of the diesel that takes place at nearly stoichiometric
conditions, irrespective of the overall lean mixture, due to spontaneous ignition and ame propagation. Flame then propagates
throughout the entire mixture, but the early stage of the combustion process is especially important as the burned gases are compressed to a higher temperature, increasing the thermal NO
formation rate [7,52]. Additionally, the rich-fuel spray core is
responsible for the prompt NO formation, particularly as the
combustion chamber contains a sufcient amount of active radicals
formed throughout the compression stroke as a result of the preignition chemical reactions of the hydrogen fuel-air mixture. It is
well accepted that the dual-fuel engine describes its combustion
process as proceeding in three stages. Initially, around half of the
pilot will burn and entrain some gaseous fuel into an overall fuelrich process. Subsequently, the remaining pilot fuel burns and entrains an increasing amount of the primary fuel (hydrogen gas) into
its reaction zone. In the nal stage, a ame propagates throughout
the combustion chamber and consumes the remaining gaseous fuel
[31].
It may be noted that in case of variable speed automotive engines, NO emission level depends on all the aforementioned three
parameters ((i) combustion temperature, (ii) oxygen concentration,
and (iii) reaction time). But, in case of constant speed stationary
engines, as the reaction time available for fuel-air mixture could be
constant, the reaction time may not be the inuencing parameter
for the emission formation. With hydrogen addition in to the engine under dual-fuel mode, oxygen content in the combustion
chamber may decrease slightly due to decrease in volumetric efciency of the engine. Decrease in oxygen content in the combustion
chamber would result in NO emission reduction in the engine [52].
However, the NO emission increased drastically with increase in
hydrogen energy shares, which indicate that slight reduction in the
oxygen content has insignicant effect on the emission formation.

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

Hence, it could be concluded that the drastic increase of the NO


emission in the engine is mainly due to temperature effect. Other
two aspects; reaction time and oxygen content are not contributing
towards drastic increment in the emission from the hydrogen dualfuel engine. Based on the above assessment, the NO emission formation via thermal route was focused in the study. Thermal NO
emission was predicted with two zone model using extended Zeldovich mechanism (Fig. 5).
It could be observed from Fig. 5 that the NO emission forms
about at 364 CA in the engine under base diesel mode at 100%
load. However, the starting point of the NO emission formation
advanced slightly with increasing amount of hydrogen substitution.
For example, the emission formation started at about 360 CA and
358 CA with 16.7% and 18.8% hydrogen energy shares respectively
in the engine under dual-fuel mode. The main reason for early
formation of the emission could be due to advanced start of combustion with increase in hydrogen energy share. It could also be
observed from Fig. 5 that the entire emission formation occurred
during early part of the combustion process i.e., from about 360 CA
to about 378 CA, which clearly indicates the premixed combustion
is a major inuencing phase for the emission formation. The earlier
experimental investigation on hydrogen dual-fuel engine supports
this fact of NO emission formation in premixed combustion phase
[23]. They stated that the increase in premixed combustion phase
would enhance the NO formation rate signicantly. Simulation
results of Maghbouli et al. also evident that the entire NO emission
formed between 358 CA to 385 CA during premixed combustion
phase in dual-fuel combustion mode [26]. It is evident that high
temperature of combusted products during this premixed

481

combustion phase are favourable for high level of NO emission


formation. Masood and Ishrat carried out both experimental and
computational work to understand the effect of in-cylinder temperature on NO emission in a hydrogen dual-fuel engine [38]. They
reported the increase in NO mole fraction from 0.012 at the incylinder temperature of 2000 K to 0.028 at 3000 K [38]. Cumulative NO emission in the engine under dual-fuel mode at 100% load
increased signicantly with increase in hydrogen energy share as
seen in Fig. 5. The cumulative emission increased from 863 ppm
with base diesel mode (0% hydrogen energy share) to 1307 ppm
with 18.8% hydrogen energy share under dual-fuel mode. The
similar trend of increase in cumulative NO emission was reported
by Maghbouli et al. in a dual-fuel engine [26].
The NO emission results obtained with two zone model were
validated with the experimental test data (CI engine: 7.4 kW rated
power at 1500 rpm) as shown in Fig. 6. The emission increased with
increase in hydrogen energy share signicantly. For example, the
emission increased from 759 ppm with 0% hydrogen energy share
(diesel mode) to 1148 ppm with 18.8% hydrogen energy share
(dual-fuel mode). NO emission predicted with localized burned
zone temperature is higher than the measured NO emission at all
hydrogen energy shares. For example, at 18.8% hydrogen energy
share, the experimental value of the emission is 1148 ppm whereas
the predicted value is 1308 ppm (based on localized burned zone
temperature). This is mainly because of neglecting the reverse reactions as well as crevice volume for computation of the emission.

5.2. Nitric oxide (NO) emission prediction using CFD simulation


5.2.1. Mesh independence and convergence analysis
Fig. 7 shows CFD simulation of in-cylinder pressure variation
inside the combustion chamber of the hydrogen dual-fuel CI engine
for the considered mesh types (Mesh type1, type 2 and type 3). It
could be clearly seen from the results that in-cylinder peak pressure increased with increasing mesh density up to a certain extent
and then remains unchanged. For example, the in-cylinder peak
pressure with mesh type 2 (mesh density: 1.6 node spacing) is
higher (79 bar) than the pressure (76 bar) with mesh type 1 (mesh
density: 2 node spacing) as shown in Fig. 7. However, the incylinder peak pressures with Mesh type 3 and Mesh type 4 are
same about 83 bar and there is no signicant difference between
the accuracies of the in-cylinder pressure proles obtained with
these both mesh types (Fig. 7). Hence, in view of less computational

Fig. 5. NO formation rate and cumulative NO emission in the engine using two zone
model.

Fig. 6. Comparison of NO emission with experimental and two zone model results.

482

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

Fig. 7. In-cylinder pressure curves with respect to degree crank angle for different
meshes.

cost for analysing less mesh density domain and comparably high
accuracy, the Mesh type 3 was selected for further simulation work.
Fig. 8 shows the convergence criteria for different parameters with
respect to time steps. It is observed from Fig. 8 that all parameters
including energy, velocity, k and epsilon satised the convergence
criteria (Table 12) for every time step.
5.2.2. Validation of CFD results with experimental results
Simulated in-cylinder pressure and in-cylinder temperature
results obtained from the CFD analysis on Mesh type 3 were validated with the experimental results for the same engine operating
conditions. Fig. 9 shows validation of CFD simulation results of incylinder pressure and in-cylinder temperature with the experimental results for base diesel and dual-fuel modes at 100% load. Incylinder peak pressure was slightly over predicted (particularly in
premixed combustion phase) and subsequently this error was
amplied in case of in-cylinder temperature. It could also be clearly
seen that start of combustion (SOCcfd) was advanced in CFD simulation as compared to experimental results (SOCexp). For example,
at 16.7% hydrogen energy share SOCcfd occurred at 2 CA after TDC
whereas SOCexp occurred at 4 CA after TDC. This over prediction of
in-cylinder pressure and in-cylinder temperature could be due to

Fig. 9. Validation of in-cylinder pressure and in-cylinder temperature using CFD


simulation.

Fig. 8. Convergence history of different parameters with respect to time steps.

neglecting crevice volumes, blow-by losses, mismatching of


compression ratio and uncertainty associated with equivalence
ratio. Similar type of trends was reported by Maghbouli et al. in
their CFD simulation work in a dual-fuel engine [26]. For example,
they found the simulated in-cylinder pressure was about 76 bar
whereas the experimental in-cylinder pressure was about 68 bar in
the engine. They stated the main reason for over prediction was due
to high autoignition heat release followed by faster combustion
reaction rates [26]. Liu and Karim also expressed the similar results
of higher in-cylinder peak pressure value (about 78 bar) with CFD
simulation than the experimental in-cylinder pressure about 74 bar
in a dual-fuel engine [53]. The variations in the in-cylinder pressure

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

and in-cylinder temperature are affected by the total kinetic energy


variations following piston movement, with a slight pressure difference existing between the swirl and the main combustion
chambers [53]. Belal et al. also found over predication of in-cylinder
peak pressure by about 6% in a dual-fuel engine combustion
simulation [54]. Knop et al. stated that the slight pressure overestimation before combustion for the direct injection engine was
mainly associated with an uncertainty on the compression ratio
[36]. The calculations were performed with the theoretical
compression ratio while the real compression ratio appears to be
slightly lower [36]. Simulation work carried out by Azimov et al.
revealed that equivalence ratio also plays a signicant role for
better prediction of in-cylinder pressure during the combustion
[34]. They found the best CFD simulation results as compared with
experimental data in the range of equivalence ratios below about
0.8 in a dual-fuel engine. As the equivalence ratio increased above
0.8, the results deviated from the experiment data [34]. In the
present CFD simulation also, a sudden rise in the pressure and incylinder temperature could be observed at top dead centre as
shown in Fig. 9. High percentage of error band about 15%e23% was
observed between CFD simulation and the experimental results.

483

effect of turbulence on NO emission formation needs to be carried


out.
Fig. 11 depicts the clear distinction among the spatial distribution of in-cylinder temperature with base diesel mode and dualfuel modes. In case of dual-fuel mode, rapid combustion of fuel
air charge could be clearly seen in Fig. 11 due to presence of
hydrogen gaseous fuel. This is one of the signicant benets with
hydrogen dual-fuel operation. Once localized in-cylinder temperature was determined, the NO emission formation was computed
using Zeldovich mechanism in the CFD simulation. Thermal NO and
prompt NO models were used in the study for accurate prediction
of the emission formation in the engine. Figs. 12 and 13 show
spatial distribution of NO emission and the corresponding incylinder temperature with respect to degree CA for base diesel
and dual-fuel modes. It could be observed from Figs. 12 and 13 that
the degree CA at which the emission formation started is earlier in
case of dual-fuel mode than base diesel mode. The rise in high incylinder temperature during premixed combustion phase would
lead to increase in NO emission formation. In a typical diesel

5.2.3. Effect of hydrogen addition on turbulence and NO emission in


the engine
Fig. 10 shows CFD simulation of turbulent kinetic energy variation inside the combustion chamber for different hydrogen energy
shares in the dual fuel engine. Turbulence increased slightly with
increase hydrogen energy share in the engine. For example, the
peak turbulent kinetic energy increased from 90 m2/s2 with 14.5%
hydrogen energy share to 120 m2/s2 with 18.8% hydrogen energy
share (Fig. 10). Liu and Karim also reported the simulation results of
increase in turbulent kinetic energy from 125 m2/s2 to 170 m2/s2
during combustion (from 0 CA to 40 CA after TDC) inside combustion chamber of a dual-fuel engine [53]. It could also be
observed from the results that turbulent kinetic energy increases
signicantly with proceeding in combustion process. Other than
the three inuencing parameters (in-cylinder temperature, oxygen
content and reaction time) which were discussed in earlier Section
5.1, the NO emission formation may also depend on the turbulence
inside the combustion chamber. Increase in turbulence may affect
the NO emission formation as high turbulence may lead to increase
in the in-cylinder temperature with enhanced air-fuel mixing rates
inside the combustion chamber. However, the detailed study on the

Fig. 10. Turbulent kinetic energy variation with respect to degree CA using CFD
simulation.

Fig. 11. Contours of in-cylinder temperature with respect to degree crank angles.

484

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

Fig. 12. Contours of NO emission and the corresponding in-cylinder temperature at different degrees crank angle of the engine under base diesel mode.

engine, the premixed combustion phase generally starts at crank


angle nearly TDC (360 CA) and remains till 10e15 CA after TDC. It
can be observed from the in-cylinder temperature contours in
Figs. 12 and 13 that the temperature rise due to premixed combustion occurs earlier (advanced) crank angle in case of dual fuel
engine than base diesel engine operation. Flame speed of
hydrogen-air premixed charge during combustion is also faster in
the dual fuel operation than base diesel operation. These are some
reasons observed from the CFD results for higher NO emission
formation with dual fuel engine as compared to conventional diesel

engine. The emission formation started at 363 CA in case of base


diesel mode (Fig. 12) whereas it advanced to 360.8 CA with dualfuel mode (16.7% hydrogen energy share) (Fig. 13). The reason
could be due to faster reaction rates and higher in-cylinder temperature with hydrogen addition in case of dual-fuel mode. With
hydrogen addition in the engine, concentration of OH free radicals
may increase signicantly that would enhance the chemical reaction rate and subsequently the NO emission formation rate. The
studies conducted by Masood et al. also revealed the substantial
increase in combustion velocity from 385 m/s with 20% hydrogen

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

485

Fig. 13. Contours of NO emission and the corresponding in-cylinder temperature at different degrees crank angle under dual-fuel mode with 16.7% H2 energy share.

energy share to 670 m/s with 80% hydrogen energy share in a


hydrogen based dual-fuel engine [37]. At 363 CA, contours of NO
emission are widespread in case of dual-fuel mode as compared to
diesel mode (Fig. 13). Even at 380 after TDC, the localized incylinder peak temperature is higher (about 2490 K) with dual
fuel mode than base diesel engine (about 2450 K) as shown in
Figs. 12 and 13. A notable conclusion is emerged from the CFD results that the in-cylinder temperature during premixed combustion

is higher with dual fuel mode than base diesel engine operation
resulting in higher NO emission. It could also be observed from the
results that the predicted (simulated) temperature is higher than
the actual temperature (calculated from measured in-cylinder
pressure data) during early part of combustion (i.e., premixed
combustion phase). This high temperature leads to prediction of
higher NO emission formation rate than the actual. This was one of
the major setbacks experienced during simulation work. However,

486

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

in the present study, an attempt has been made to work out the
methodology for NO emission formation, which is a major
pollutant in case of hydrogen based dual-fuel engines. As the NO
emission increases exponentially with increase in in-cylinder
temperature, operation of the engine at low in-cylinder temperatures would give the substantial benet of low NO emission in
hydrogen dual-fuel engines. For example, Chintala and Subramanian utilized water injection strategy for reduction of NOx
emission in a hydrogen dual-fuel engine, and achieved 37% emission reduction with 270 g/kWh water addition [3,6,8].
5.3. Comparison of nitric oxide (NO) emission with theoretical and
experimental results
Fig. 14 shows the comparison of NO emission results with
theoretical modelling (two zone model and CFD simulation) and
experimental tests in the engine under dual fuel mode at 100% load
and 1500 rpm. It could be observed from the gure that the
emission predicted trend is similar to the experimental trend with
the error band of 15%e23%. However, the emission determined
using two zone model or CFD simulation is higher than the
experimental values. This may be due to higher simulated localized
temperatures and ignoring blowby and crevice losses. At 16.7%
hydrogen energy share, the emission values with experimental test,
two zone model, and CFD simulation are 914 ppm, 1208 ppm, and
1382 ppm respectively. These results are in agreement with the
earlier study, that the NO emission was higher (about 680 ppm)
with CFD analysis than the experimental results (470 ppm) at 20%
hydrogen energy share in a dual-fuel engine [37]. In the present
study, a methodology for predicting the NO emission in a hydrogen
dual-fuel engine was explored qualitatively, however high percentage of error in the emission prediction is a major setback for the
theoretical study. In case of CFD simulation, unavailability of exact
chemical reactions mechanism with diesel and hydrogen fuels is a
major drawback. In addition to this, neglecting crevice volume and
blow-by effects for the simulation are the other major reasons for
obtaining high percentage of error with respect to experimental
results.

and experimental studies were conducted to get a deeper insight


on the effect of localized in-cylinder temperature on the NO
emission formation in a dual-fuel engine (7.4 kW rated power at
1500 rpm). The following conclusions are drawn based on the
experimental and simulation results.
Localized in-cylinder temperature in burned zone is signicantly higher than that of the global in-cylinder temperature
which is calculated from the measured in-cylinder pressure
data. At 16.7% hydrogen energy share, the in-cylinder peak
temperature in burned zone is about 2403 K whereas the global
peak temperature is about 1840 K.
NO emission formation rate increased steeply during premixed
combustion phase. The NO emission mainly formed during
363e375 CA. The starting point (degree CA) of emission formation advanced with increase in hydrogen energy share in the
engine under dual-fuel mode.
Higher in-cylinder pressure and in-cylinder temperatures were
predicted with CFD simulation as compared to the measured
data (in-cylinder pressure) during early part of combustion (i.e.,
premixed combustion phase). This high temperature leads to
prediction of higher NO emission formation rate than the actual.
This was one of the major drawbacks experienced during
simulation work as blow-by and crevice losses were neglected
for simulation. However, the emission results are in agreement
with the experimental results qualitatively with the error band
of 15%e23%.
The present study results could be used for development of
hydrogen dual fuel engines with low NO emission. Theoretical/
CFD methodology used in the study could also be applicable for
dual fuel engines with other gaseous fuels including natural gas
and biogas.

Appendix

6. Conclusions

Uncertainty associated with hydrogen energy share.


Hydrogen energy share under dual-fuel mode is given in Eq.
(A.1)

High amount of oxides of nitrogen (NO) emission is one of the


major problems in hydrogen dual-fuel engines. Both theoretical

Hydrogen energy share HES

_ H2 CVH2
m
_ diesel CVdiesel
_
mH2 CVH2 m
(A.1)

Mass ow rate of hydrogen 0.0333 g/s.


Mass ow rate of diesel 0.4 g/s.
%Error
associated
with
hydrogen
ow
rate
measurement 0.124.
%Error associated with diesel ow rate measurement 0.178.
With substitution of values for CV and mf of fuels Eq. (A.1) could
be written as Eqs. (A.2-A.3).

HES

120E03  mf ;H2
120E03  mf ;H2 43E03  mf;diesel

(A.2)

120E03  0:0333E  03
 100
120E03  0:0333E  03 43E03  0:4E  03
18:8%

HES

(A.3)

Fig. 14. Comparison of NO emission among experimental, two zone model and CFD
simulation results.

With reference to Eq. (A.2), the uncertainty associated with


hydrogen energy share (HES) can be expressed as given in Eqs.
(A.4eA.10)

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

HES fmH2 ; mdiesel


"

DHES

(A.4)
[13]

vHES
_ H2
Dm
_ H2
vm

2

vHES
_ diesel
Dm
_ diesel
vm

_ diesel
vHES
5160E06  m

_ H2
vm
_
_
120E03  m
43E03  m
H2

diesel

2 #0:5
(A.5)

[14]

[15]
2

[16]

5160E06  0:4E  03

[17]

120E03  0:0333E  03 43E03  0:4E  032

4594:11

[18]

(A.6)
_ hydrogen
5160E06  m
vHES

_ diesel
vm
_ H2 43E03  m
_ diesel 2
120E03  m

[19]

[20]

5160E06  0:0333E  03
120E03  0:0333E  03 43E03  0:4E  032

[21]

382:46
(A.7)

[22]

_ H2 0:03330:124=100 4:13E  05
Dm

(A.8)

[23]

_ diesel 0:40:178=100 0:000712


Dm

(A.9)

[24]

DHES

0:5

2
4594:115  4:13E05  382:46  0:0007122

0:33%

[25]

[26]

(A.10)
Actual hydrogen energy share (HES) with inclusion of
uncertainty 18.8 0.33.

[27]
[28]

References
[29]
[1] Subramanian KA, Singal SK, Saxena M, Singhal S, et al. Utilization of liquid
biofuels in automotive diesel engines: an Indian perspective. Biomass Bioenergy 2005;29(1):65e72.
[2] Babu MKG, Subramanian KA. Alternative transportation fuels, utilisation in
combustion engines. Taylor & Francis Group: CRC Press; 2013.
[3] Chintala V, Subramanian KA. Experimental investigation of hydrogen energy
share improvement in a compression ignition engine using water injection
and compression ratio reduction. Energy Convers Manag 2016;108:106e19.
[4] Subramanian, K.A. and V. Chintala, Reduction of GHGs emissions in a biodiesel
fueled diesel engine using hydrogen in ASME 2013 internal combustion engine fall technical conference. 2013, paper No. ICEF2013e19133, doi:10.1115/
ICEF2013-19133: [Dearborn, Michigan].
[5] Chintala V, Subramanian KA. Experimental investigations on effect of different
compression ratios on enhancement of maximum hydrogen energy share in a
compression ignition engine under dual-fuel mode. Energy 2015;87:448e62.
[6] Chintala V, Subramanian KA. An effort to enhance hydrogen energy share in a
compression ignition engine under dual-fuel mode using low temperature
combustion strategies. Appl Energy 2015;146:174e83.
[7] Heywood JB. Internal combustion engines fundamentals. New York: McGrawHill, Inc; 1988.
[8] Chintala V, Subramanian KA. Hydrogen energy share improvement along with
NOx (oxides of nitrogen) emission reduction in a hydrogen dual-fuel
compression ignition engine using water injection. Energy Convers Manag
2014;83:249e59.
[9] Dhole AE, Yarasu RB, Lata DB, Priyam A. Effect on performance and emissions
of a dual-fuel diesel engine using hydrogen and producer gas as secondary
fuels. Int J Hydrogen Energy 2014;39(15):8087e97.
[10] Sukjit E, Herreros JM, Dearn KD, Tsolakis A, Theinnoi K, et al. Effect of
hydrogen on butanolebiodiesel blends in compression ignition engines. Int J
Hydrogen Energy 2013;38(3):1624e35.
[11] White CM, Steeper RR, Lutz AE. The hydrogen-fueled internal combustion
engine: a technical review. Int J Hydrogen Energy 2006;31(10):1292e305.
[12] Wagemakers AMLM, Leermakers CAJ. Review on the effects of dual-fuel

[30]

[31]
[32]

[33]

[34]

[35]
[36]

[37]

[38]

[39]

487

operation, using diesel and gaseous fuels, on emissions and performance.


SAE International; 2012. SAE No. 2012-01-0869.
Christodoulou F, Megaritis A. Experimental investigation of the effects of
separate hydrogen and nitrogen addition on the emissions and combustion of
a diesel engine. Int J Hydrogen Energy 2013;38(24):10126e40.
SinghYadav V, Soni SL, Sharma D. Performance and emission studies of direct
injection C.I. engine in duel fuel mode (hydrogen-diesel) with EGR. Int J
Hydrogen Energy 2012;37(4):3807e17.
se H, Ciniviz M. An experimental investigation of effect on diesel engine
Ko
performance and exhaust emissions of addition at dual-fuel mode of
hydrogen. Fuel Process Technol 2013;114:26e34.
Liew C, Li H, Gatts T, Liu S, Xu S, Rapp B, et al. An experimental investigation of
exhaust emissions of a 1999 Cummins ISM370 diesel engine supplemented
with H2. Int J Engine Res 2012;13(2):116e29.
Liew C, Li H, Liu S, Besch MC, Ralston B, Clark N, et al. Exhaust emissions of a
H2-enriched heavy-duty diesel engine equipped with cooled EGR and variable
geometry turbocharger. Fuel 2012;91(1):155e63.
Banerjee R, Bose PK. An experimental investigation on the potential of
hydrogen in the reduction of the emission characteristics of an existing fourstroke single-cylinder diesel engine operating under EGR. Int J Green Energy
2011;9(1):84e110.
Liu S, Li H, Liew C, Gatts T, Wayne S, Shade B, et al. An experimental investigation of NO2 emission characteristics of a heavy-duty H2-diesel dual-fuel
engine. Int J Hydrogen Energy 2011;36(18):12015e24.
Sahin Z, Durgun O. Multi-zone combustion modeling for the prediction of
diesel engine cycles and engine performance parameters. Appl Therm Eng
2008;28(17e18):2245e56.
Li X, Zhang A. Numerical simulation research of NOx emissions characteristic
of vehicle engine fueled with LPG. In: International conference on Electrical
and Control Engineering (ICECE); 2010.
Saravanan S, Nagarajan G, Anand S, Sampath S, et al. Correlation for thermal
NOx formation in compression ignition (CI) engine fuelled with diesel and
biodiesel. Energy 2012;42(1):401e10.
Chintala V, Subramanian KA. Experimental investigation on effect of enhanced
premixed charge on combustion characteristics of a direct injection diesel
engine. Int J Adv Eng Sci Appl Math 2014;6(1e2):3e16.
Wilhelmsson C, Tunestal P, Johansson B, Widd A, Johansson R. A physical twozone NOx model intended for embedded implementation. 2009. 2009-011509.
Rakopoulos CD, Antonopoulos KA, Rakopoulos DC. Multi-zone modeling of
Diesel engine fuel spray development with vegetable oil, bio-diesel or Diesel
fuels. Energy Convers Manag 2006;47(11e12):1550e73.
Maghbouli A, Saray RK, Shafee S, Ghafouri J, et al. Numerical study of combustion and emission characteristics of dual-fuel engines using 3D-CFD
models coupled with chemical kinetics. Fuel 2013;106:98e105.
Kumar S, Kumar Chauhan M, Varun. Numerical modeling of compression
ignition engine: a review. Renew Sustain Energy Rev 2013;19:517e30.
Peiyong N, Wang X, Wang Z, Mao G, Wei S, et al. In: Numerical modeling of
soot and NOx emissions of a diesel/methanol dual-fuel engine. In International conference on computer distributed control and intelligent environmental monitering (CDCIEM); 2011.
Mattarelli E, Rinaldini CA, Golovitchev VI. CFD-3D analysis of a light duty dualfuel (Diesel/Natural gas) combustion engine. Energy Procedia 2014;45:
929e37.
Miao H, Milton B. Numerical simulation of the gas/diesel dual-fuel engine incylinder combustion process. Numer Heat Transf Part A Appl 2005;47(6):
523e47.
Clarke A, Stewart J. A three-zone heat-release rate model for dual-fuel combustion. Proc Inst Mech Eng Part C J Mech Eng Sci 2010;224(11):2423e34.
Gharehghani A, Jazayeri A, Mirsalim M, Ghanbari M, et al. Numerical and
experimental investigation on performance of dual-fuel D87 engine. In: ASME
2010 international mechanical engineering congress & exposition,
IMECE2010; 2010 [Vancouver, British Columbia, Canada].
Hussain Shaik Magbul, Kumar BSP, Reddy KVK. CFD analysis of combustion
and emissions to study the effect of compression ratio and biogas substitution
in a diesel engine with experimental verication. Int J Eng Sci Technol 2012;4:
473e92.
Azimov U, Okuno M, Tsuboi K, Kawahara N, Tomita E, et al. Multidimensional
CFD simulation of syngas combustion in a micro-pilot-ignited dual-fuel engine using a constructed chemical kinetics mechanism. Int J Hydrogen Energy
2011;36(21):13793e807.
Rao V, Honnery D. A comparison of two NOx prediction schemes for use in
diesel engine thermodynamic modelling. Fuel 2013;107:662e70.
Knop V, Benkenida A, Jay S, Colin O, et al. Modelling of combustion and nitrogen oxide formation in hydrogen-fuelled internal combustion engines
within a 3D CFD code. Int J Hydrogen Energy 2008;33(19):5083e97.
Masood M, Ishrat M, Reddy A. Computational combustion and emission
analysis of hydrogenediesel blends with experimental verication. Int J
Hydrogen Energy 2007;32(13):2539e47.
Masood M, Ishrat MM. Computer simulation of hydrogenediesel dual-fuel
exhaust gas emissions with experimental verication. Fuel 2008;87(7):
1372e8.
Bajpai S, Sahoo PK, Das LM. Feasibility of blending karanja vegetable oil in
petro-diesel and utilization in a direct injection diesel engine. Fuel
2009;88(4):705e11.

488

V. Chintala, K.A. Subramanian / Energy 116 (2016) 470e488

[40] High speed diesel specications. India: Indian Oil Corporation Ltd. (IOCL);
2005.
[41] National hydrogen energy road map - 2006 (Abridged version, 2007). National
Hydrogen Energy Board, Ministry of New and Renewable Energy Government
of India; 2007.
[42] Das LM, Gulati R, Gupta PK. A comparative evaluation of the performance
characteristics of a spark ignition engine using hydrogen and compressed
natural gas as alternative fuels. Int J Hydrogen Energy 2000;25(8):783e93.
[43] Chintala V, Subramanian KA. A CFD (computational uid dynamics) study for
optimization of gas injector orientation for performance improvement of a
dual-fuel diesel engine. Energy 2013;57:709e21.
[44] Subramanian KA, Lahane S. Comparative evaluations of injection and spray
characteristics of a diesel engine using karanja biodiesel-diesel blends. Int J
Energy Res 2013;37(6):582e97.
[45] Turns SR. An introduction to combustion. second ed. Singapore: McGraw-Hill
Companies; 2000.
[46] ANSYS FLUENT theory guide, release 14.0. ANSYS, Inc.; 2011.
[47] Beerer DJ, McDonell VG. Autoignition of hydrogen and air inside a continuous
ow reactor with application to lean premixed combustion. J Eng Gas Turbines Power 2008;130(5):1e8.
[48] Fleck J, Griebel P, Steinberg AM, Stohr M, Aigner M, et al. Autoignition limits of
hydrogen at relevant reheat combustor operating conditions. J Eng Gas Turbines Power 2012;134(4):041502.
[49] Zhou JH, Cheung CS, Zhao WZ, Leung CW, et al. Dieselehydrogen dual-fuel
combustion and its impact on unregulated gaseous emissions and particulate emissions under different engine loads and engine speeds. Energy
2016;94:110e23.
[50] Yadav VS, Soni SL, Sharma D. Engine performance of optimized hydrogenfueled direct injection engine. Energy 2014;65:116e22.
[51] Ishida M, Yamamoto S, Ueki H, Sakaguchi D, et al. Remarkable improvement
of NOxePM trade-off in a diesel engine by means of bioethanol and EGR.
Energy 2010;35(12):4572e81.
[52] Abdelaal MM, Rabee BA, Hegab AH. Effect of adding oxygen to the intake air
on a dual-fuel engine performance, emissions, and knock tendency. Energy
2013;61:612e20.
[53] Liu C, Karim GA. Three-dimensional computational uid simulation of diesel
and dual-fuel engine combustion. J Eng Gas Turbines Power 2009;131(1).
012804: 1e9.
[54] Belal TM. Investigating Diesel Engine Performance and Emissions Using CFD.

Energy Power Eng 2013;05(02):171e80.

Abbreviations
[N2], [NO], [H2O], [O2], [O], [OH]: Nitrogen, nitric oxide, water, oxygen and hydroxyl
concentrations, mol/cm3
BDC: Bottom dead centre
bz: Burned zone
CA: Crank angle
Cp: Specic heat at constant pressure, kJ/kg-K
d(NO)/dt: Mass reaction rate of NO, g/cycle
d[NO]/dt: Molar reaction rate of NO, mol/cm3-s
H: Hydrogen atom/radical
h: Specic enthalpy, kJ/kg
H2: Hydrogen gas
_ Mass ow rate of uids/species, kg/s
m:
N: Nitrogen atom
n: Number of moles
N2: Nitrogen
NO: Nitric oxide
NOx: Oxides of nitrogen
: Degree crank angle
O: Oxygen atom
O2: Oxygen
OH: Hydroxyl radical
p: In-cylinder pressure, N/m2
Q: Heat energy/heat transfer/heat release, kW
R: Characteristic gas constant, kJ/kg-K
T: Temperature, K
TDC: Top dead centre
U: Internal energy, kW
ubz: Unburned zone
V: Cylinder instantaneous volume, m3
Y: Specic heat ratio
SOCcfd: Start of combustion with CFD simulation
SOCexp: Start of combustion with experimental tests

You might also like