You are on page 1of 61

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

net/publication/310048389

Simulation of a spark ignited hydrogen engine for minimization of NOx


emissions

Article  in  International Journal of Hydrogen Energy · November 2016


DOI: 10.1016/j.ijhydene.2016.10.074

CITATIONS READS
4 483

2 authors:

Pranav Kherdekar Divesh Bhatia


Indian Institute of Technology Delhi Indian Institute of Technology Delhi
1 PUBLICATION   4 CITATIONS    26 PUBLICATIONS   321 CITATIONS   

SEE PROFILE SEE PROFILE

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

Development of methods for simultaneous continuous monitoring of mercury, arsenic and selenium species in coal gasification process View project

All content following this page was uploaded by Pranav Kherdekar on 29 November 2017.

The user has requested enhancement of the downloaded file.


This is an accepted manuscript version for an article published in the
International Journal of Hydrogen Energy.

Copyright to the final published article belongs to Elsevier.

If you wish to cite/view the actual publication, please use the following reference
and DOI link:

Kherdekar PV, Bhatia D, Simulation of a spark ignited hydrogen engine for


minimization of NOx emissions, International Journal of Hydrogen Energy
(2016)

DOI: http://dx.doi.org/10.1016/j.ijhydene.2016.10.074
*Manuscript
Click here to view linked References

Simulation of a spark ignited Hydrogen engine for minimization of NOx emissions

Pranav V. Kherdekar1, Divesh Bhatia2*


1,2
Department of Chemical Engineering, Indian Institute of Technology Delhi, Hauz Khas, New Delhi,
INDIA - 110016

ABSTRACT

A global kinetics based model is developed for a spark ignited single cylinder H2 engine for

prediction of NO emissions, using mass and energy conservation equations. Based on literature

experimental data, a transient spark energy delivery model is developed. A comprehensive heat

transfer model is used which considers the heat losses by a combination of forced convection,

conduction and natural convection using extended surfaces. It is shown that the major resistance

to heat transfer in our setup is because of convective heat transfer from the cylinder to the

coolant, with the in-cylinder heat transfer playing a relatively minor role. The model predicts the

temporal variation of pressure and temperature profiles during the various strokes of the engine

operation and their effect on the transient NO concentrations. It is shown that the exhaust NO

concentrations depend not only on the peak in-cylinder temperatures, but also on the temporal

variation of the in-cylinder temperature. A thermodynamically consistent rate expression for NO

formation is used, which predicts a maximum in the in-cylinder NO concentration with time due

to the decrease in temperatures during the power stroke and the reversible nature of NO

formation. Using a kinetics based approach, the effect of various operating variables such as

engine speed, equivalence ratio, compression ratio and spark advance on the NO emissions is

predicted by the model, with results consistent with reported trends. It is shown that the NO

emissions go through a maximum with engine speed and the range of engine speeds for which
2

the NO emissions increase is broader for high equivalence ratios. The maximum in NO

emissions with equivalence ratio and the increase in NO emissions with spark advance and

compression ratio is predicted, which is in agreement with the literature. It is found that the

equivalence ratio strongly affects the peak in-cylinder temperature, whereas the engine speed

governs the amount of time available for NO formation. Consistent with reported data, the

phenomenon of back-fire is predicted at high equivalence ratios. It is found that NO emissions

are sensitive to compression ratio at high values of the compression ratio. Optimum values of the

equivalence ratio and compression ratio are suggested to optimize the power output and NO

emissions.

Keywords: NOx emissions; H2 engine; Emission modeling; Spark delivery; Kinetics; Simulation

*Corresponding author. Email: dbhatia@chemical.iitd.ac.in


3

1. INTRODUCTION

Although the idea of using hydrogen as an energy source was acknowledged decades ago, the

advent of hydrogen as an engine fuel and the subsequent evolution of hydrogen engines can be

attributed to the advantages of hydrogen over conventional fossil fuels. Considering the gradual

depletion of conventional hydrocarbon based fuels and the increasing stringency of emission

norms, the use of hydrogen as an engine fuel is in focus because of its conditional renewability,

economic viability, and low emission of pollutants. Hydrogen is found to be superior to fossil

fuels also on account of faster flame propagation and production of a steady flame unlike heavy

fossil fuels [1]. Also, Veziroglu and Sahin [2], in their study regarding economics of hydrogen

usage as an alternative fuel concluded that hydrogen has an unparalleled economic advantage

over fossil fuels.

Hydrogen combustion in internal combustion engines is free from emission of toxic pollutants

like carbon monoxide, oxides of sulfur, particulate matter, etc. The major product formed during

hydrogen combustion is water, in contrast to hydrocarbon based fuels where the combustion

products contain significant amounts of carbon dioxide, a greenhouse gas. Hydrogen is a cleaner

fuel as compared to carbon based fuels but it is not totally devoid of harmful emissions. At high

temperatures attained in the combustion chamber during combustion of hydrogen, thermal

fixation of air results in significant formation of thermal NOx. The rate of NOx formation is

highly sensitive to cylinder temperature at high temperatures exceeding 1600 K [3]. Various

physical and operating parameters such as engine heat loss, spark timing, combustion mixture

composition and compression ratio affect the cylinder temperature and also play an important

role in NOx formation in the combustion chamber [3].


4

Many efforts have been made to understand the nature of these intertwined aspects of IC engines

on their performance and emissions. Different phenomenological and multi-dimensional models

have been developed in various studies for combustion, heat transfer, flow and emissions for IC

engines [4]. However, despite availability of abundant kinetic data on hydrogen combustion and

nitrogen oxides’ formation, many simulation works on hydrogen engines have considered

empirical modeling of hydrogen combustion [5-7]. An empirical model used for predicting

conversion of fuel as a function of only the crank angle is the cosine-burning law [8]. Another

semi-empirical model widely used is the Wiebe’s function that also approximates the fuel

conversion as a function of only the crank angle, but correlated by fuel-specific constants to fit

the burning rate profile [9-10]. Furthermore, the same values of the constants as those used in

modeling gasoline combustion have been used in some hydrogen engine simulations [11]. This

compromises with the accuracy of prediction of performance and estimation of in-cylinder

temperature and pressure and subsequently with the accuracy of prediction of NOx emissions.

This is due to the significant difference in the burn velocities of the two fuels and the assumption

that crank angle is the sole parameter determining the fuel conversion in a hydrogen engine [12 -

13].

Various models have expanded the traditional view towards hydrogen engines of being

mechanical devices to a more comprehensive one. Several studies of zero-dimensional, multi-

zone and multidimensional modeling of hydrogen engines have been reported. Fagelson et al.

[14] reported one of the earliest works on thermodynamic modeling of a hydrogen fuelled SI

engine. They developed a quasi-dimensional two-zone model using a semi-empirical turbulent

flame expression to predict power output and NO emissions. The model was later applied by

Kumar et al. [15] to investigate the performance and NO emissions of a supercharged hydrogen
5

engine. A quasi-dimensional model with a laminar burning velocity correlation was developed

by Verhelst and Sierens [16] for studying the pressure and temperature in a SI hydrogen engine.

D’Errico et al. [17] developed a quasi-dimensional multi-zone combustion model for a hydrogen

engine with a cryogenic port injection system. The work included 1-D simulation of intake and

exhaust processes. They investigated and validated the effect of engine speed, air-fuel ratio and

spark advance on the exhaust NO emissions

Some studies have considered the effect of turbulence on hydrogen combustion and NOx

formation in the engines [18-19]. Multidimensional models have been employed for hydrogen

engines by means of development of CFD simulation codes. Johnson [20] studied the effect of

ignition timing and engine speed on NO emissions at a constant equivalence ratio using Kiva-3V

engine simulation code. Adgulkar et al. [21] simulated a hydrogen engine using AVL-FIRE

simulation package.

With the availability of detailed kinetic mechanisms like those provided by The Gas Research

Institute (GRI 3.0), models based on detailed kinetics have been developed. Al-Baghdadi et al.

[22] modeled performance of a single cylinder hydrogen engine by considering detailed kinetics

for modeling hydrogen combustion. Safari et al. [23] investigated the effect of equivalence ratio,

spark advance and compression ratio on NO emissions with application of detailed kinetics.

They also studied the effect of lean burn and EGR techniques on exhaust NO emissions.

Increase in modeling intricateness however results into increased computational efforts whereas

application of simple global mechanisms with the assumption of uniform spatial distribution in

an engine cylinder has shown satisfactory results [24]. Marinov et al. [25] validated the global

kinetics proposed by Marinov et al. [26] for hydrogen combustion in hydrogen fuelled internal
6

combustion engines. On the other hand, the validity of Zeldovich mechanism for NOx formation

has been substantiated by its extensive use in major works associated with thermal NOx

formation [27].

This work focuses on developing a phenomenological model for predicting NO emissions for a

single cylinder hydrogen engine with a minimum usage of empirical expressions. The model is

based on fundamental conservation principles and kinetics of hydrogen combustion and NO

formation. Hydrogen as an engine fuel has unique transport properties especially in terms of

inter-specie diffusion and in terms of heat transfer characteristics. With diffusion coefficient at

NTP around 12 times higher than gasoline and thermal conductivity at NTP approximately 16

times higher than gasoline, hydrogen also burns with a high flame speed, causing rapid

combustion [28]. Hence, we neglected the minor diffusional resistances and propagation of flame

front and the single cylinder hydrogen engine was assumed to be a perfectly mixed variable

volume reactor. The intake and exhaust flow rates were modeled so as to depict the pressure-

volume characteristics similar to those observed in real-time operation [7]. The spark energy

delivery was modeled as one composed of a combination of high energy release rate period and a

constant low energy release rate period. For computing heat losses from the engine cylinder, heat

transfer from cylinder wall and extended surfaces by natural convection was considered and the

effect of heat losses on the exhaust NO concentration was studied. The model comprising of

ordinary differential equations for species and energy balances was used to generate profiles of

the in-cylinder temperature, pressure and NO concentration and subsequently the exhaust NO

concentrations. The results of the simulations were compared with data available from various

works on the hydrogen engine.


7

2. DESCRIPTION OF SETUP AND INTERPRETATION OF THE PROCESS

The operational and constructional details of the hydrogen engine under consideration have been

obtained from a single cylinder engine at IIT Delhi. It is a four-stroke, spark-ignited single

cylinder internal combustion engine in which pure H2 is injected using fuel injectors. The engine

cylinder (combustion chamber) is made up of cast-iron wall, and a piston reciprocates inside the

cylinder covering a displacement volume of 394 cc. The cylinder is provided with an intake

valve, an exhaust valve and a spark plug. Atmospheric air is used as coolant for the cylinder and

annular fins are provided at the periphery of the cylinder in order to ensure efficient cooling of

gases within the chamber. A schematic of the engine is shown in Fig. 1. The geometric

parameters, heat transfer properties and operating variables are given in Table 1.

Fig. 1 Schematic of the simulated engine


8

Table 1 Geometric parameters, heat transfer properties and operating variables of the simulated engine

Property Symbol Value Units


Geometry
Bore B 86 10-3 m
Length of connecting rod Lr 86 10-3 m
Stroke s 68 10-3 m
Crank radius ac 34 10-3 m
Cylinder wall thickness 6.5 10-3 m
Displacement volume Vd 394 10-6 m3
Clearance volume Vc 44 10-6 m3
Exhaust valve diameter Dexh 15.7 10-3 m
Number of fins Nfin 13 -
-3
Fin thickness xfin 3 10 m
Length of fin Lfin 16 10-3 m
Effectiveness of fin εfin 3.8
Compression ratio rc 9

Induction and outflow


Volumetric efficiency of ηvol 0.65 -
induction
Exhaust manifold pressure Pout 1.115 105 Pa
Ambient Pressure Patm 1.01 105 Pa
Discharge coefficient of Cd 0.6 -
exhaust valve
Heat transfer properties
Material Cast iron
Ambient temperature Tatm 298 K
Specific heat ratio γ 1.33 -
Emissivity of gas ε 0.6 -
Spark Energy Delivery Ed 25 J/cycle
9

The overall process of operation of the hydrogen engine resembles the Otto Cycle and is shown

in Fig. 2. The first stroke is an intake process in which the intake valve opens and air-fuel

mixture enters the combustion chamber on account of pressure deficit created due to the motion

of piston from top dead center (TDC) to bottom dead center (BDC). This can be assumed as a

gaseous mixing process at constant pressure [7]. In the second stroke, both valves are closed and

the gaseous mixture is compressed which results in an increase in its temperature. In our model,

this process is considered to be the adiabatic compression of an ideal gas with the assumption of

negligible heat loss to the surroundings [7]. In the third stroke, the mixture already at a high

temperature is ignited by means of an electric spark. Owing to high inter-specie diffusion

coefficients and the fast propagation of hydrogen flame, this process can be assumed to be a

chemical reaction in a well-mixed batch reactor, occurring in a single zone [28]. The fourth

stroke finally consists of pumping out the gaseous mixture from the cylinder with driving force

provided by the motion of piston towards TDC to reduce the cylinder volume. The resistance to

this outflow of exhaust gases is offered by the constriction of exhaust valve. This final stroke can

hence be assumed as flow of a high pressure gas through a nozzle to the exhaust manifold [29].

3. MODELING APPROACH

The species balance equations and the energy balance equation were developed on lines similar

to those employed by Kumar et al. [30] in modeling of a gasoline fueled spark ignition engine.

The phenomenological model was further expanded to simulate the hydrogen engine under

consideration. It was reported by Karim [28] that the effects of swirl are insignificant in a

hydrogen engine due to the high molecular diffusivity and the high burning velocity of H2.

Hence, mixing effects were not considered for the engine and perfect mixing of cylinder contents

was assumed. The time variation of spark release rate was included in the model and a detailed
10

model of the heat losses by natural convection was developed. In our work, we modeled the four

strokes of the engine as interdependent processes. In the initial simulations, we used both the

Vander Wall’s equations and ideal gas law for the calculations and the results were observed to

be similar. Due to the close proximity of gas pressure predicted by the Vander Wall’s equation

and the ideal gas law under the typical process conditions, the gaseous mixture was assumed to

behave as an ideal gas for the sake of simplicity.

(a) (b) (c) (d)

Fig. 2 Four-stroke combustion cycle

3.1 Species Balance

Pure hydrogen is injected into the combustion chamber along with atmospheric air. The air is

considered to be composed of nitrogen and oxygen only, and three reactant species (H2, O2, and

N2) are initially fed to the reactor, which undergo the following global reactions to form water

and NO:

Reaction 1

Reaction 2
11

Of these, the forward reaction (2) is undesirable.

Since it is assumed that the contents of the reactor are uniformly mixed, volume averaged

concentrations were considered in the model. For any ith specie amongst the five species

involved in the reactions, the rate of change of concentration can be written as [30]

(1)

The term Ri in Eq. (1) represents the net rate of generation/consumption of component i due to

the two individual reactions (j=1 to 2). Mathematically, for any component i,

(2)

where rj is the rate of reaction j and is the stoichiometric coefficient of component i in

reaction j.

3.2 Energy Balance

Energy balances and subsequently the equation of temperature consider the rate of energy input

to the cylinder by intake gases and spark, enthalpy change due to reactions, rate of energy

flowing out with the products, heat rejected to the coolant and thermodynamic effects of volume

change under closed conditions. Eq. (3) gives the rate of temperature change with respect to time

[30]:

(3)
12

3.3 Reaction Kinetics

We have used global kinetics as a trade-off between empirical modeling and detailed micro-

kinetics. For H2 oxidation, we used a global kinetic model proposed by Marinov et al. [26]. The

model was tested for hydrogen combustion in an internal combustion engine and was found valid

against a wide range of laminar speeds for lean hydrogen-fuelled operation [25]. The rate of

hydrogen combustion can be expressed as

(4)

For NO formation, we employed a global kinetic expression that considers Zeldovich mechanism

[9,31]. Marinov et al. [25] also corroborated the Zeldovich mechanism by concluding that major

NO formation in a hydrogen engine is due to the interaction between N2 molecules and O

radicals. The rate expression for NO formation can be expressed as

(5)

where, Keq is the equilibrium constant for the reversible reaction which is

calculated from the Gibbs free energy as follows:

(6)

Thus, a thermodynamically consistent rate expression is used in the current work.


13

3.4 Flow rates

Opening and closing of the valves affects the flow rate terms occurring in Eqs. (1) and (3). In the

intake stroke and exhaust stroke, the chamber behaves as an open system allowing inflow and

outflow of gases, respectively. In the compression stroke and power stroke, the chamber behaves

as a closed system wherein adiabatic compression and batch conversion of contained gases,

respectively, takes place.

3.4.1 Intake Stroke

During the intake stroke, air and fuel that are premixed in the intake manifold enter the

combustion chamber due to low pressure created by motion of the piston from TDC to BDC.

The intake valve is assumed to open at a crank angle of 0o and close at 180o [30]. The rate of

intake of air is assumed to be proportional to the rate of change of cylinder volume, and is given

as

(7)

The constant of proportionality is volumetric efficiency of induction, which is defined as the

ratio of volumetric flow rate of air into the cylinder to the rate of displacement of volume by the

piston [9]. Although models are available to predict volumetric efficiency, experimentally

determined values are more reliable. It was found that for a single cylinder variable compression

SI engine similar to the one under consideration, the volumetric efficiency varies between 62%

and 70% with an average value of 65% [32]. We performed simulations with varying values of

volumetric efficiency between 62% and 70% and observed that varying the volumetric efficiency

did not have a significant impact on the exhaust NO concentrations. Hence, we used a constant
14

value of volumetric efficiency in the present work and is given in Table 1. A limitation of using

Eq. (7) is that if the rate of change of cylinder volume is negative, the predicted flow rate of air

would not be physically realistic. This will happen if the intake valve opens before the piston

reaches TDC or it remains open even after the piston reaches BDC.

Flow rates of fuel and air are related by equivalence ratio, which is defined as the ratio of actual

fuel-air ratio and stoichiometric fuel-air ratio in Eq. (8) as

(8)

Specification of equivalence ratio represents the excess/deficit of air relative to the

stoichiometric conditions. Fuel flow rate as a function of flow rate of air and equivalence ratio is

expressed in Eq. (9) as

(9)

where is the mole fraction of oxygen in atmospheric air and is the stoichiometric

oxygen-to-fuel ratio on molar basis.

Assuming a negligible temperature change due to mixing, the total flow rate to the chamber is

obtained by addition of air flow rate and the fuel flow rate, as

(10)

The mole fraction of each component in the air-fuel mixture is obtained by

(11)
15

Assuming ideal gas law to be applicable, the inlet concentration and molar flow rate of each

component is computed as

(12)

(13)

3.4.2 Exhaust Stroke

The exhaust valve is assumed to open at a crank angle of 540o and close at 720o [30]. The rate of

removal of exhaust gases from the combustion chamber is given by the discharge equation

obtained from Bernoulli obstruction theory as [29]

(14)

The molar flow rate of each component is computed using Eq. (15), assuming that the

concentration of the components being removed is equal to their in-cylinder concentration.

(15)

The inlet and outlet molar flow rates defined in Section (3.4.1) and (3.4.2), respectively, and

used in Eqs. (1) and (3) are summarized in Table 2 for various strokes.

3.5 Calculation of instantaneous volume and rate of change of volume of the cylinder

Due to the reciprocating motion of piston, the volume of cylinder changes as the piston vacillates

between TDC and BDC. The rate of change of volume with respect to time needs to be

calculated since it is required in Eqs. (1), (3) and (7). The instantaneous volume can be expressed

as a function of crank angle (θ) by slider crank mechanism as [9]


16

(16)

The rate of change of instantaneous volume can be found out by differentiating Eq. (16) with

respect to crank angle and substituting the expression for crank angle as a function of time. It is

given by Eq. (17) as follows:

(17)

Table 2 Intake and exhaust flow rates for different strokes

Stroke Fi(in) Fi(out)

Intake Stroke 0

Compression Stroke 0 0

Power Stroke 0 0

Exhaust Stroke 0

3.6 Spark ignition

The initiation of hydrogen combustion reaction takes place due to a spark. Hence, spark delivery

to compressed gases needs to be modeled. Also, ignition timing has a significant effect on engine

performance. For gasoline engines, it is well studied that spark initiates the process of oxidation

of fuel but complete combustion of fuel takes a finite time [9]. If the spark is delivered prior to

piston reaching the TDC, more fuel conversion is completed before TDC is reached. The

combined effect of an increase in pressure and temperature due to combustion initiation and

compression results in a relatively higher peak cylinder temperature, pressure and power output.
17

This technique of delivering spark before the piston reaches TDC is called spark advancing and

is used widely for increasing the power output from gasoline engines. However, as spark

advancing results in elevated temperatures, it is expected to increase NO emissions from the

engine. Our model considers the effect of angle of spark advancing on NO emissions.

Accordingly, the angle of spark advance before top dead center (BTDC) is specified (θadvance) and

the angle of spark start (θspark,start) is computed as given in Eq. (18).

(18)

The angle at which spark delivery is terminated (θspark,stop) is computed by specifying the duration

of spark delivery as

(19)

We simulated the transient profile of energy delivery rate by a spark such that it is similar to that

observed experimentally by Kleinhenz et al. [33]. Figure 3 shows a typical simulated spark

energy delivery curve for an engine speed of 2400 RPM and an energy release of 25 J per spark.

Energy delivery by spark refers to the release of energy by the spark into the engine cylinder

during spark ignition.

The spark energy delivery starts at tspark,start and terminates at tspark,stop. For first one fourth of the

period of spark, the energy delivery is assumed to be an intense energy release spike and for the

rest of the period, constant energy delivery is assumed. It is assumed that the rate of energy

delivery by the spark in the constant energy release period is one-fourth of the maximum energy

delivery rate during the spike. The rates of increase and decrease in the rate of energy delivery

(slope of ascent and descent of spike) are computed by equating the area under the spark energy

delivery rate curve to the total energy delivered per spark (Ed).
18

The time at which the energy release spike starts is equal to the time of start of spark delivery.

The spike is followed by a constant energy delivery rate which starts at tspike,stop and is calculated

as

(20)

Fig. 3 Typical spark delivery curve

The time at which the rate of energy delivery is maximum is denoted by tspike in Fig. 3 and is

given as

(21)

The maximum rate of energy delivery at this time (tspike) is given by


19

(22)

For the time range between tspark,start and tspike, Eq. (23) gives the instantaneous rate of energy

released by the spark as

(23)

Between tspike and tspike,stop, the instantaneous rate of energy released by the spark is given by Eq.

(24) as

(24)

For the remaining spark delivery period, a constant energy release rate was considered and is

given by Eq. (25) as

(25)

3.7 Heat transfer

Parameters including but not limited to exothermicity of reactions, enthalpy of reactants and

products, work done by piston and compression of the gases significantly alter the cylinder

temperature. As NO formation is strongly dependent on the in-cylinder temperature, heat loss to

the surroundings is of foremost consideration in modeling of NO emissions from the hydrogen

engine.

Heat transfer takes place between in-cylinder gaseous mixture and the inner cylinder wall via

convection and radiation. The heat transfer further takes place through cylinder wall by means of

conduction and from the outer surface of the cylinder to the air (coolant) by means of natural
20

convection, with the latter being aided by peripheral annular fins. Fig. 4 shows a schematic of the

heat-transfer geometry and the expected trends in the temperature profile.

Fig. 4 Heat transfer from engine (axisymmetric view)

The wall temperature on gas side (Twg) is computed by equating the rate of heat transfer from the

in-cylinder gases to the inner surface, the heat transfer rate within the wall by conduction and the

heat transfer rate from the outer surface to the ambient air. This results in the following algebraic

equation which can be solved for Twg, if the values of in-cylinder temperature and ambient

temperature are known:

(26)
21

After determination of the inner wall temperature, the temperature of wall on the outer surface

(Twc) is calculated by equating the rate of heat conduction through the wall and the rate of heat

transfer by convection from wall to the ambient air, as

(27)

Solution of Eq. (26) and use of Eq. (27) requires specification of the thermal conductivity of cast

iron, in-cylinder heat transfer coefficient (hcg) and heat transfer coefficient on coolant side of the

wall (hcc). Heat transfer coefficient within the engine cylinder has been widely computed by

using Woschni’s correlation in major engine simulation works on hydrogen engine [6,11,13,34-

35]. Demuynck et al. [36] showed that the heat transfer models of Annand and Woschni were not

suitable for use in H2 engines. Wei et al. [37] showed that Woschni’s model underpredicts the in-

cylinder heat transfer coefficient for a H2 engine by a factor of two. Hence, the in-cylinder heat

transfer coefficient is predicted by Woschni’s correlation modified with a correction factor

(Fc=2) given by Eq. (28) as

(28)

Here, the characteristic velocity ‘w’ considers the effect of mean piston speed as well as the

effect of combustion, and is given as

(29)

The motored pressure (Pm) is computed by Eq. (30) assuming engine operation without

combustion of fuel, i.e., for a process consisting of adiabatic expansion and compression without

reaction.
22

(30)

The constants C1 and C2, as defined by Woschni are given in Table 3.

Table 3 Parameters for Woschni correlation

Crank angle range C1 C2

θ=0 to π 6.18 0

θ= 3π to 4π

θ= π to 2π 2.28 0

θ= 2π to 3π 2.28 3.24 10-3

In the system under consideration, heat transfer from cylinder wall to the ambient air takes place

by natural convection. Cooling of cylinder contents is enhanced by extended surfaces (annular

fins) present at the outer periphery of the cylinder wall (Fig. 4). Heat is transferred from cylinder

wall to the ambient air through annular fins and the cylinder wall directly exposed to the ambient

air. The coolant side heat transfer hence needs to consider heat losses from the exposed surface

area of cylinder wall as well as from the annular fins. The value of coolant side heat transfer

coefficient at different temperatures and effectiveness of fin is obtained by steady state

simulation of the engine heat transfer in COMSOL®. Geometry similar to the engine cylinder is

constructed and a range of temperatures from 300 K to 2400 K are imposed on inner wall of the

cylinder. Thermal conductivity of cast iron is formulated as a polynomial function of temperature

[38]:
23

(31)

Convective heat flux is obtained by surface integration over the exposed wall of the cylinder and

the exposed area of fins. This is then used to calculate the coolant side heat transfer coefficient

and effectiveness of the fin. The heat transfer coefficient at various temperatures is given in Fig.

5. It is observed that the heat transfer coefficient increases with an increase in the inner wall

temperature. This is because heat transfer coefficient for natural convection is strongly

dependent on the Grashof number which in turn depends on the difference in wall and ambient

temperature.

The outer surface area of cylinder exposed to the coolant is given by Eq. (32) as

(32)

Fig. 5. Variation of heat transfer coefficient on coolant side (h cc) with cylinder wall temperature (Twc)
24

The area of cylinder wall that forms base for the fins and hence is not exposed to the coolant is

given as

(33)

The total rate of heat transfer to the coolant is given as

(34)

εfin represents the effectiveness of annular fins and is defined as the ratio of actual heat loss from

the fins to the heat loss from the base area if the fins were not employed. For the temperature

range between 300 K and 2400 K, the average fin effectiveness was calculated to be 3.8 with a

standard deviation of 0.18.

3.8 Solution technique and description of the solver

The energy balance equation and the species balance equations for H2, O2, N2, H2O and NO were

solved simultaneously using ODE23s solver in MATLAB®. The variables were non-

dimensionalized in the simulation in order to avoid huge differences in the orders of magnitude

of the variables being solved. A stiff solver was used because of the high reaction rates and

extremely fast change in the variables with respect to time. The time scales associated with the

simulation were of the order of a few milliseconds.

Since the heat transfer from the in-cylinder gases to the coolant takes place by a combination of

forced convection, natural convection, radiation and conduction, we performed a sensitivity

analysis to study the phenomenon which resulted in the maximum resistance to heat transfer. It

was found that natural convection at the outer surface of the cylinder effectively governed the

rate of heat transfer. Hence, using a correction factor of 2 in the Woschni’s correlation to
25

calculate the internal heat transfer coefficient did not have a significant impact on the overall

heat transfer coefficient, and hence on the predicted heat transfer rates. Around 0.1-2%

difference was observed in the maximum in-cylinder temperature. However, since NO

concentration is strongly dependent on temperature, 0.05-13% difference was observed in the

exhaust NO concentration on revising the in-cylinder heat transfer model. Had the same cylinder

been cooled with a flowing coolant, we expect that the increase in overall heat transfer

coefficient would have been higher and the accuracy of the correlation for the internal heat

transfer coefficient would have been more crucial for getting accurate predictions.

4. RESULTS AND DISCUSSION

4.1 Transient profiles

The ordinary differential equations were solved for an engine speed of 2400 RPM, equivalence

ratio of 0.5, compression ratio of 9 and no advancing of spark. The other parameters used for the

simulation are given in Table 1. These are the base values of the parameters and will be used in

other simulations as well, unless otherwise specified. From the simulation run, the transient

profiles of in-cylinder pressure, temperature and NO concentrations were obtained.

Fig. 6 (a) shows the simulated in-cylinder pressure profile for the first few combustion cycles. It

is observed that a periodic steady state is achieved, i.e., the pressure profile does not change for

subsequent cycles. The peak pressure obtained during the power stroke is around 40 bar, whereas

the pressure obtained at the end of exhaust stroke is around 1 bar. Fig. 6 (b) shows the variation

of in-cylinder pressure and motored pressure as a function of crank angle when a periodic steady

state is achieved, i.e., when the profiles of in-cylinder pressure, temperature and NO

concentration do not change for subsequent cycles. During the power stroke, a significant
26

difference between the in-cylinder pressure under actual operation and the in-cylinder pressure

under motored condition indicates that combustion is taking place effectively. It is also observed

that the motored pressure and operating cylinder pressure values are almost the same before

commencement of the spark, and cylinder pressure departs from motored pressure after

commencement of spark. Hence, during the compression stroke, the pressure can be reliably

calculated at any given cylinder volume by assuming an adiabatic process (PVγ=constant), and

considering that P=Pin at VR=Vc+Vd. Though pressure does not appear directly in the kinetic

expression for NO formation, it appears in the piston work term in Eq. (3), thus affecting the in-

cylinder temperature that directly appears in the rate law for NO formation. To check the validity

of our simulation results, we also performed simulations using the geometry of cylinder given by

Jilakara et al. [35]. It was observed that the model-predicted variation of in-cylinder pressure

with crank angle was in agreement with their experimental data trends.

45

40

35
In-Cylinder Pressure (bar)

30

25

20

15

10

0
0 0.1 0.2 0.3 0.4 0.5
Time (s)
27

(a)

In-cylinder pressure Motored pressure


45
Intake Compression Power Exhaust
40 Stroke stroke stroke Stroke

35

30
Pressure (bar)

25

20

15

10

0
0 180 360 540 720
Crank angle (degrees)

(b)

2500

2000
In-cylinder Temperature (0C)

1500

1000

500

0
0 0.1 0.2 0.3 0.4 0.5
Time (seconds)
28

(c)

2500
Intake Compression Power Exhaust
stroke stroke stroke stroke

2000
In-Cylinder temperature (0C)

1500

1000

500

0
0 180 360 540 720
Crank Angle (degrees)

(d)

1000

900
In-cylinder NO concentration (ppm)

800

700

600

500

400

300

200

100

0
0 0.1 0.2 0.3 0.4 0.5
Time (seconds)
29

(e)

1200
Intake Compression Power Exhaust
stroke stroke stroke stroke
1000
In-cylinder NO concentration (ppm)

800

600

400

200

0
0 180 360 540 720
Crank Angle (degrees)

(f)

Fig. 6 Simulation results at 2400 RPM and φ=0.5 (a) Transient variation of in-cylinder pressure (b) Variation

of in-cylinder pressure and motored pressure with crank angle (c) Transient variation of in- cylinder

temperature (d) Variation of in-cylinder temperature with crank angle (e) Transient variation of in- cylinder

NO concentration (f) Variation of in-cylinder NO concentration with crank angle

Fig. 6 (c) shows the simulated in-cylinder temperature profile for the first few combustion

cycles. It is observed that the in-cylinder temperatures reach more than 2000 oC, which is

responsible for NO formation. Fig. 6 (d) represents the simulated in-cylinder temperature

variation with respect to crank angle after attainment of a periodic steady state. With the

attainment of a steady state, the intake stroke is characterized by dilution of accumulated high

temperature gases from the previous cycle by low temperature gaseous-mixture entering the

cylinder, hence reducing the in-cylinder temperature as seen in Fig. 6 (d). From Fig. 6 (b), it can
30

be seen that the intake process takes place at a nearly constant pressure. The intake stroke is

followed by compression of gases, leading to an increase in the temperature of gases till the

onset of spark, as is seen in Fig. 6 (d). As soon as the spark is given, the initial phase of power

stroke is characterized by an abrupt increase in the in-cylinder temperature due to the heat

released by fuel combustion. The immediate and drastic temperature increase after spark delivery

represents very fast conversion of hydrogen. This is in agreement with experimental observations

but in contrast to the burned mass fraction model by Wiebe [25,39]. The later phase of power

stroke is characterized by a decrease in the in-cylinder temperature due to work done by the

gases and also due to heat losses from the cylinder. Finally, the exhaust stroke is characterized by

a further gradual decrease in the in-cylinder temperature.

Fig. 6 (e) shows the simulated profile of in-cylinder NO concentration for the first few

combustion cycles. In the first few cycles, complete combustion of fuel does not take place

resulting in lower in-cylinder temperatures. Hence, low NO concentrations are attained during

the first few cycles until a periodic steady state is attained. Fig. 6 (f) represents the simulated

variation of in-cylinder NO concentration with respect to crank angle once a periodic steady state

is achieved. It was discussed earlier that the intake-stroke is characterized by dilution of residual

gases accumulated from the previous cycle. This causes a drop in the in-cylinder NO

concentration which is evident from Fig. 6 (f). The NO concentration attained up to the end of

intake stroke remains constant during the compression stroke. This is because the temperature

attained at the end of the compression stroke is not high enough for significant NO formation

rates to be reached. A sharp temperature increase after spark delivery is accompanied by a drastic

increase in the in-cylinder NO concentration. This confirms that NO formation is not sensitive to

temperature at the low temperatures attained during compression stroke but very sensitive to
31

higher temperatures attained due to the combustion of fuel during the power stroke. This is

attributable to the high activation energy of NO formation.

As the complete combustion cycle takes place in a few milliseconds, the instantaneous in-

cylinder NO concentration is not measured experimentally. Also, the in-cylinder temperature is

estimated from other measured variables. Accordingly, experimental data for their variation with

crank angle is not available. However, the trends in the variation of in-cylinder NO concentration

were found to be in agreement with those reported by other IC-engine simulation works

[30,32,40].

4.2 Parametric Sensitivity

It was reported by various research groups that the major factors affecting NOx emissions from

hydrogen engine are engine speed, equivalence ratio, spark timing, heat transfer provisions and

the compression ratio [3,22,41-42]. To elucidate the role of these parameters, we conducted

different simulation runs by varying one parameter at a time, while keeping the other parameters

constant. Since NO emissions take place only during the exhaust stroke, the exhaust NO

concentration was taken to be the same as the in-cylinder NO concentration during the exhaust

stroke.

4.2.1 Effect of engine speed

Engine speed is a primary parameter that affects NOx emissions from a hydrogen engine [22].

Fig. 7 (a) represents the simulated NO emissions and the maximum in-cylinder temperature as a

function of engine speed at an equivalence ratio of 0.5. It is observed that in the range of normal

idling speeds (600-1000 rpm), NO emissions increase with respect to engine speed, reach a

maximum at the engine speed of 1200 RPM and decrease thereafter. A possible explanation for
32

this behavior can be proposed as follows: The time available per cycle is inversely proportional

to the engine speed. At very low engine speeds, time available for heat losses to the ambient air

is more and hence in-cylinder temperatures remain low due to efficient heat transfer, leading to

low NO emissions. As the engine speed is increased, the time available for heat transfer

decreases, leading to higher in-cylinder temperatures and hence higher NO emissions. However,

beyond a certain engine speed, this ‘high temperature period’ available for NO formation starts

decreasing, thus providing less time for NO to be formed. Thus, despite the monotonic increase

in maximum in-cylinder temperature with the engine speed, the exhaust NO concentration does

not show a similar trend. These trends of temperature with engine speed are in agreement with

the data reported by Rahman et al. [6].

Exhaust NOx concentration Maximum in-cylinder Temperature

240.00 2100

2080

Maximum Cylinder Temperature (0C)


220.00
Exhaust NO concentration (ppm)

2060
200.00
2040
180.00
2020
160.00
2000
140.00
1980

120.00 1960

100.00 1940
600 1200 1800 2400 3000 3600
Engine Speed (RPM)

(a)
33

600
hcc=f(Tw-Tc)

500 hcc=[f(Tw-Tc)]/2
Exhaust NO concentration (ppm)

400

300

200

100

0
600 1000 1400 1800 2200 2600 3000 3400
Engine speed (RPM)

(b)

Exhaust NO concentration (ppm) maximum in-cylinder temperature

2800 2420

2750

Maximum in-cylinder temperature (0C)


2400
Exhaust NO concentration (ppm)

2700
2380
2650
2360
2600
2340
2550
2320
2500
2300
2450

2400 2280

2350 2260
600 1200 1800 2400 3000 3600
Engine Speed (RPM)

(c)
34

φ=0.65 φ=0.5

16000 1200
Intake Compression Power Exhaust

In-cylinder NO concentration (ppm) for φ=0.5


In-cylinder NO concentration (ppm) for φ=0.65
14000 stroke stroke stroke stroke
1000
12000
800
10000

8000 600

6000
400
4000
200
2000

0 0
0 180 360 540 720
Crank angle (degrees)

(d)

Figure 7 (a) Variation of exhaust NO concentration and maximum in-cylinder temperature with engine speed

for φ=0.5 (b) Effect of coolant side heat transfer coefficient on exhaust NO emissions for φ=0.5 (c) Variation

of exhaust NO concentration and maximum in-cylinder temperature with engine speed for φ=0.65 (d) In-

cylinder NO concentration profile for φ=0.65 and φ=0.5 for 2400 rpm

The sensitivity of NO emissions to heat losses was found by performing a simulation with the

heat transfer coefficient on the coolant side put as half of the value calculated for the base

simulations. The variation of exhaust NO concentrations with engine speed is plotted in Fig. 7

(b) for the two heat transfer coefficients. It is observed that the decrease in the coolant side heat

transfer coefficient results in an increase in the exhaust NO concentration at any engine speed.

This is because with decrease in the coolant side heat transfer coefficient, heat losses to the

ambient air decrease, leading to attainment of higher in-cylinder temperatures and higher NO

emissions.
35

Fig. 7 (c) shows the variation of exhaust NO concentration and maximum in-cylinder

temperature with engine speed at an equivalence ratio of 0.65. It is observed that the nature of

variation of the exhaust NO concentration with engine speed is different from that observed for

the equivalence ratio of 0.5 (Fig. 7 (a)). For a given engine speed, the NO emissions for an

equivalence ratio of 0.65 are higher than for 0.5. This is because of the higher values of peak in-

cylinder temperatures obtained for high values of equivalence ratios, as observed by comparison

of Figs. 7 (a) and 7 (c).

It is also observed that for the high equivalence ratio of 0.65, the range of engine speed over

which NO emissions increase is broader than for the case with equivalence ratio of 0.5. As

discussed earlier, the decrease in NO emissions with engine speed takes place due to a decrease

in the time available for NO formation. For the case of high equivalence ratios, the in-cylinder

temperatures are higher than the case with lower equivalence ratios (Figs. 7 (a) and 7 (c)). It was

also discussed in Section 4.1 that NO formation depends strongly on the in-cylinder temperature

at high temperatures. Thus, for high equivalence ratios, the effect of increasing temperature

dominates over the effect of reduced time period available for NO formation. This results in an

increase in NO emissions with an increase in the engine speed over a broad range of engine

speeds.

The in-cylinder NO concentration is plotted in Fig. 7 (d) for equivalence ratios of 0.5 and 0.65,

once a periodic steady state is achieved. For an equivalence ratio of 0.65, it is observed that the

in-cylinder NO concentration goes through a maximum, close to the initial phase of the power

stroke. This suggests that at very high temperatures attained at the onset of power stroke, the rate

of the backward reaction becomes substantial. However, as the in-cylinder

temperature decreases due to further motion of piston towards BDC, the rate of backward
36

reaction decreases and becomes negligible. Further decrease in the in-cylinder temperature with

the motion of piston towards BDC causes the rate of forward reaction also to

be negligible. Thus, after ~40oCA ATDC in the power stroke, the in-cylinder NO concentration

remains constant. The maximum in the in-cylinder NO concentration is not observed for the

lower equivalence ratio (φ=0.5). This is because the temperatures attained are not sufficient

enough to initiate the backward reaction.

4.2.2 Effect of equivalence ratio

From Fig. 7 (a) and Fig. 7 (c), it can be inferred that the equivalence ratio significantly affects

the exhaust NO concentration. This is in agreement with experimental findings [41]. Fig. 8 (a)

shows the variation of exhaust NO concentration with equivalence ratio at different engine

speeds. As the equivalence ratio increases, the amount of fuel intake during the intake stroke

increases. Burning of more fuel leads to attainment of higher gas temperatures that leads to

higher NO emissions. With a further increase in the equivalence ratio, the NO emissions

decrease and the value of equivalence ratio at which NO emissions are the highest depends on

the engine speed. A maximum in the NOx concentration was reported at an equivalence ratio of

0.75 [43], which is close to the value of ~0.7 reported in the present study. Similarly,

Subramanian et al. [44] reported a maximum in NOx concentration at an equivalence ratio

between 0.7 and 0.9. Also, Jabbr et al. [45] performed simulations on a single cylinder H2 engine

using 3D AVL and predicted a maximum in the NOx concentration at an equivalence ratio of 0.8.

Thus, our predictions are in-line with the experimental data reported elsewhere. The trend of NO

concentration with respect to equivalence ratio is explained next by considering the effect of

temperature on the reversibility of NO formation reaction.


37

Figs. 8 (b) and 8 (c) show the variation of in-cylinder temperature and NO concentration with

respect to crank angle during the power stroke for equivalence ratios of 0.65 and 0.75,

respectively, at an engine speed of 2400 RPM. It can be seen that on increasing the equivalence

ratio to 0.75, maximum in-cylinder temperatures above 2500 oC are attained. This results in peak

in-cylinder NO concentrations higher than those obtained for an equivalence ratio of 0.65.

However, the in-cylinder NO concentration at the end of power stroke is lower for the

equivalence ratio of 0.75 as compared to the equivalence ratio of 0.65. This is because of the

reverse reaction , which takes place at high temperatures obtained at high

equivalence ratios, resulting in low in-cylinder NO concentrations. This shows that the NOx

emissions are dependent not only on the peak in-cylinder temperatures, but are influenced also

by the transient temperature variation within the cylinder. The trends of variation of in-cylinder

NO concentration with crank angle at different equivalence ratios are in agreement with the

results by Safari et al. [23].


38

3000
1200 RPM

Exhaust NO concentration (ppm) 2500 2400 RPM


3600 RPM
2000

1500

1000

500

0
0.3 0.4 0.5 0.6 0.7 0.8 0.9
Equivalence Ratio

(a)
In-cylinder temperature In-cylinder NO concentration
3000 18000

16000

In-cylinder NO concentration (ppm)


2500
14000
In-cylinder temperature (0C)

2000 12000

10000
1500
8000

1000 6000

4000
500
2000

0 0
360 380 400 420 440 460 480 500 520 540
Crank Angle (degrees)

(b)
39

In-cylinder temperature In-cylinder NO concentration


3000 18000

16000

In-cylinder NO concentration (ppm)


2500
14000
In-cylinder temperature (0C)

12000
2000
10000

1500 8000

6000
1000
4000

2000
500
0

0 -2000
360 380 400 420 440 460 480 500 520 540
Crank Angle (degrees)

(c)
40

(d )

8000

7000
In-Cylinder H2 concentration (ppm)

6000

5000

4000

3000

2000

1000

0
0 180 360 540 720
Crank Angle (degrees)
41

(e)

6500

6000

5500
Indicated Power (W)

5000

4500

4000

3500

3000

2500

2000
0.3 0.4 0.5 0.6 0.7 0.8
Equivalence Ratio

(f)
Figure 8 (a) Variation of exhaust NO concentration with equivalence ratio at different engine speeds (b)

Variation of in-cylinder temperature and in-cylinder NO concentration with crank angle at φ=0.65 (c)

Variation of in-cylinder temperature and in-cylinder NO concentration with crank angle at φ=0.75 (d)

Contour showing effect of equivalence ratio and engine speed (e) Variation of in-cylinder H2 concentration

with crank angle at φ=0.9 (f) Variation of indicated power with equivalence ratio

Fig. 8 (d) shows the simultaneous effect of equivalence ratio and engine speed on exhaust NO

concentration. It can be interpreted that for equivalence ratios less than 0.65, the effect of

equivalence ratio on exhaust NO concentration is stronger than the effect of engine speed as the

contour lines are almost parallel to the engine speed axis. For equivalence ratios greater than

0.65, the exhaust NO concentration varies substantially with engine speed.

Fig. 8 (e) shows the in-cylinder hydrogen concentration as a function of crank angle at φ=0.9. It

is observed that the H2 concentration decreases rapidly at the onset of the intake stroke (θ=0).
42

This is because of the high rates of H2 oxidation reaction at the high temperatures attained for

high equivalence ratios. Thus, the model predicts ignition at the start of intake stroke, referred to

as back-fire, rendering one unable to use such high equivalence ratios [46]. Even though not

shown here, the phenomenon of back-fire was predicted for equivalence ratios greater than or

equal to 0.9. The occurrence of back-fire is attributable to the lower ignition energy of hydrogen.

From simulations, we also found that for equivalence ratios below 0.3, H2 ignition and the

corresponding pressure rise was not predicted. The operable range of 0.3-0.9 beyond which

simulation exhibits no ignition and back-fire respectively is in close approximation of the

generally accepted flammability range (0.23 < φ < 1) for hydrogen combustion [47].

Fig. 8 (f) depicts the effect of equivalence ratio on indicated power at a constant speed of 2400

RPM. The indicated power increases almost linearly with equivalence ratio. On the other hand,

as observed in Fig. 8 (a), an increase in exhaust NO concentration is seen as the equivalence ratio

is increased above 0.4. Hence, use of equivalence ratio in the range of 0.5-0.55 seems to be the

most ideal with simultaneous consideration of NO emissions and engine power output. The

simulated variation of the exhaust NO concentration with equivalence ratio is in agreement with

the experimental data reported by Khajuria [32].

4.2.3 Effect of spark advancing

One of the strategies used for increasing the power output of an engine is advancing the

commencement of spark to a few degrees before the piston reaches TDC. Fig. 9 (a) shows the

effect of spark advancing on the maximum in-cylinder temperature attained at different engine

speeds, for an equivalence ratio of 0.5. It can be seen that the maximum temperature of the in-

cylinder gaseous mixture increases with spark advance.


43

2250
1200 RPM
2400 RPM
Maximum in-Cylinder temperature (0C) 2200
3600 RPM

2150

2100

2050

2000
0 5 10 15 20 25 30 35
Angle of spark advance (degrees)

(a)

1600
1200 RPM
1400
2400 RPM
Exhaust NO concentration (ppm)

1200 3600 RPM

1000

800

600

400

200

0
0 5 10 15 20 25 30 35
Angle of spark advance (degrees)

(b)
44

(c)

Fig. 9 (a) Variation of maximum in-cylinder temperature with spark advance (b) Variation of exhaust NO

concentration on spark advance (c) Contour showing effect of engine speed and spark advance on NO emissions

Fig. 9 (b) shows the exhaust NO concentration at different angles of spark advance. As NO

formation is highly sensitive to temperatures, the increase in maximum temperature attained with

advancing of spark (Fig 9 (a)) results in higher NO emissions. The trend of NO emissions with

angle of spark advance at different engine speeds is depicted in Fig. 9 (c). The near-uniform

spacing between the lines suggests that NO emissions increase uniformly with angle of spark

advance at any engine speed. The increase in NO emissions with spark advance was also

reported by Chaichan and Abass [3] and Subramanian et al. [44].


45

4.2.4 Effect of compression ratio

Apart from controllable parameters like engine speed, spark timing and equivalence ratio, a

major design aspect of hydrogen engine - the compression ratio also plays an important role in

affecting the NO emissions. Compression ratio is the ratio of the maximum volume of the engine

cylinder to its minimum volume, and represents the extent to which the mixture is being

compressed. It is expected that an increase in compression ratio will result in higher

temperatures, hence resulting in high NOx emissions [40]. This trend is valid for the range of

compression ratios in which the engine is operable. The minimum compression ratio is 7 for the

engine under consideration. At a compression ratio lower than the minimum value, the

temperature of hydrogen-air mixture is not sufficient to initiate the combustion process.

Figure 10 (a) shows the variation of indicated power with engine speed for different compression

ratios at an equivalence ratio of 0.5. It can be observed that increasing the compression ratio

from 7 to 9 at any given engine speed increases the indicated power. A gradual increase in the

indicated power is observed with a further increase in the compression ratio.

Fig. 10 (b) shows the variation of maximum in-cylinder temperature with compression ratio at

various engine speeds. It can be seen that the maximum in-cylinder temperatures increase

monotonically with an increase in the compression ratio. Thus, an increase in the compression

ratio is expected to cause an increase in the exhaust NO concentration. Fig. 10 (c) shows the

effect of compression ratio on the exhaust NO concentration. As the compression ratio is

increased from 7 to 9, the exhaust NO concentrations do not increase as significantly as they do

when the compression ratio is increased above 9. This can be explained using the trends obtained

for maximum in-cylinder temperature with a variation in compression ratio. It is observed in Fig.
46

10 (b) that the maximum in-cylinder temperature increases gradually when compression ratio is

increased from 7 to 9. This is followed by a rapid increase in the maximum in-cylinder

temperature with a further increase in the compression ratio.

Fig. 10 (d) represents the simultaneous effect of compression ratio and engine speed on the

exhaust NO concentration. It is observed that the distance between the near parallel lines

decreases for high compression ratios, indicating a rapid increase in NO emissions for high

compression ratios. As both the power output and exhaust NO concentrations increase with

increase in compression ratio, there needs to be a compromise between the power output and

exhaust NO concentration. As seen from Fig. 10 (a), the indicated power changes gradually after

reaching a compression ratio of 9. However, in the same range of compression ratios, NO

concentrations increase rapidly as observed in Fig. 10 (c) and 10 (d). This suggests that a

compression ratio of 9-9.5 is optimal for efficiently running the hydrogen engine without

significant NO emissions and without compromising with the power output.


47

6000
1200 RPM
5500
2400 RPM
Indicated Power (W) 5000 3600 RPM

4500

4000

3500

3000

2500

2000

1500
7 8 9 10 11 12
Compression ratio

(a)

2250
1200 RPM
Maximum in-cylinder temperature (0C)

2400 RPM
2200
3600 RPM

2150

2100

2050

2000

1950
7 8 9 10 11 12
Compression Ratio

(b)
48

1400
1200 RPM

Exhaust NO concentration (ppm) 1200 2400 RPM


3600 RPM
1000

800

600

400

200

0
7 8 9 10 11 12
Compression Ratio

(c)
49

(d)

Fig. 10 (a) Variation of indicated power with compression ratio (b) Variation of maximum in-cylinder

temperature with compression ratio (c) Variation of exhaust NO concentration with compression ratio (d)

Effect of compression ratio and engine speed on exhaust NO concentration

5. Conclusions

A model is developed for a spark-ignited single-cylinder hydrogen engine for prediction of NO

emissions, using mass and energy conservation equations. Global kinetics are used for modeling

of hydrogen combustion and NO formation reactions. Spark energy delivery to the engine

cylinder is modeled in line with the experimental data. A comprehensive model that considers in-

cylinder heat transfer coupled with heat transfer to the coolant using extended surfaces reduced

the use of empirical correlations in modeling of engine heat transfer. The use of a modified
50

Woschni’s correlation with a higher heat transfer coefficient for hydrogen is shown to produce a

relatively minor effect on the overall heat transfer coefficient, and hence the in-cylinder

temperature.

Effect of four operating and geometric parameters viz. engine speed, equivalence ratio, spark

timing and compression ratio on exhaust and in-cylinder NO concentration is studied. The model

predicts that the exhaust NO concentration goes through a maximum with an increase in engine

speed. For high equivalence ratios (φ>0.6), the NO concentration increases with engine speed for

a broader range of the engine speed. It is shown that the equivalence ratio strongly impacts the

maximum in-cylinder temperature while the engine speed governs the high temperature period

available for NO formation, which causes the different NO emissions behavior for various

equivalence ratios. It is shown that exhaust NO concentrations depend not only on the maximum

in-cylinder temperature but also on the temporal evolution of temperature profile during the

power stroke. Furthermore, it is also shown that at higher equivalence ratios (>~0.7), the rate of

backward reaction increases substantially, which causes a maximum in the in-

cylinder NO concentration with respect to time. The model effectively captures the

experimentally observed trend of a maximum in NO concentration with respect to equivalence

ratio and the phenomenon of back-fire at high equivalence ratios. The maximum with respect to

equivalence ratio is predicted by considering the dependence of temperature on the reversibility

of the NO formation reaction. An equivalence ratio of around 0.5 is found to be optimum with

simultaneous consideration of NO emissions and engine power output while spark advance is

found to cause an increase in the NO emissions. The increase in exhaust NO concentration with

compression ratio is predicted by the model and compression ratio in the range of 9-9.5 is

suggested for optimizing the power output and NO emissions.


51

Acknowledgements

We thank Mr. Jayakrishnan K. Unni of Indian Institute of Technology Delhi for assistance with

the construction and operational details of hydrogen fuelled single cylinder IC engine. His

expertise on H2 engines has assisted us in our research.

List of symbols and terms

Symbol Description Units

ac Crank radius m

Acyl,exposed Area of cylinder outer wall surface exposed to ambient air m2

Acyl,fin Area of cylinder outer wall surface covered with fins m2

ATDC After top dead center -

Avalve Area of exhaust valve m2

B Bore of the engine cylinder m

BDC Bottom dead center -

BTDC Before top dead center -

Cd Discharge coefficient of exhaust valve -

Ci Average concentration of specie i in the cylinder mol/m3

Ci,in Concentration of specie i at the inlet of the cylinder mol/m3

Cpi Molar heat capacity of specie i J/mol-K

C1 Woschni’s constant for the mean speed term -

C2 Woschni’s constant for the combustion term -

Dexh Exhaust valve diameter m

Ed Energy delivered to engine cylinder per spark J


52

Fc Correction factor for Woschni’s correlation -

Fi(in) Molar flow rate of specie i into the cylinder mol/s

Fi(out) Molar flow rate of specie i out of the cylinder mol/s

hcc Coolant side heat transfer coefficient W/m2K

hcg In-cylinder heat transfer coefficient W/m2K

Hi(in) Molar enthalpy of specie i at the inlet of the cylinder J/mol

Hi Molar enthalpy of specie i at the temperature of cylinder J/mol

k Thermal conductivity of cast iron W/(m-K)

Keq Equilibrium constant for NO formation reaction -

l Thickness of cylinder wall m

Lcyl Length of cylinder m

Lfin Length of a fin m

Lr Length of connecting rod m

nfuel Number of moles of fuel entering the cylinder mol

Number of moles of oxygen entering the cylinder mol

Nc Number of components involved in the process -

Nfin Number of annular fins -

NR Number of reactions -

P In-cylinder pressure Pa

Patm Ambient pressure Pa

Pin Inlet manifold pressure Pa

Pm Motored pressure Pa
53

Pout Exhaust manifold pressure Pa

Pr Reference pressure for Woschni’s correlation Pa

c Rate of heat transfer to ambient air (coolant) J/s

Qin,air Volumetric flow rate of air into the cylinder m3/s

Qin,fuel Volumetric flow rate of fuel into the cylinder m3/s

Qin,total Total volumetric flow rate into the cylinder m3/s

max Maximum rate of energy release by the spark J/s

Qout Volumetric flow rate out of the cylinder m3/s

s Instantaneous rate of energy release by the spark J/s

rc Compression ratio -

rcyl,o Outer radius of the cylinder m

rj Rate of reaction j mol/m3s

R Ratio of crank radius to length of the connecting rod -

Rg Gas constant J/mol-K

Ri Rate of consumption of component i mol/m3s

s Stroke m

Sp Mean piston speed m/s

t Time s

tspark,start Time at which spark delivery is initiated s

tspark,stop Time at which spark energy delivery is terminated s

tspike Time at which maximum rate of spark energy release is s

attained
54

tspike,stop Time at which spark energy delivery as a spike is s

terminated

T In-cylinder temperature K

Tatm Atmospheric temperature K

Tc Ambient air temperature (Same as atmospheric temperature) K

TDC Top dead center -

Tin Temperature of gases entering the engine cylinder K

Tr Reference temperature for Woschni’s correlation K

Twc Temperature of cylinder wall exposed to ambient air K

Twg Temperature of inner wall of cylinder K

Vc Clearance Volume m3

Vd Displacement volume m3

VR Instantaneous volume occupied by gases in the cylinder m3

Vr Reference volume of cylinder for Woschni’s correlation m3

w Characteristic velocity for Woschni’s correlation m/s

xfin Thickness of fin m

xi,air Mole fraction of component i in air -

xi,fuel Mole fraction of component i in fuel -

xi,in Mole fraction of component i in the air-fuel mixture at -

intake valve

Mole fraction of oxygen in air -

γ Specific heat ratio of hydrogen-air fuel mixture -

ε Emissivity of the gaseous mixture -


55

εfin Effectiveness of the fins -

ηvol Volumetric efficiency of induction -

Number of moles of oxygen required per mole of fuel for -

stoichiometric combustion

Stoichiometric coefficient of component i for reaction j -

φ Equivalence ratio -

ρ Density of air-fuel mixture in the cylinder kg/m3

σ Boltzmann’s constant Wm-2K-4

ω Angular speed of rotation of crank rad/s

θ Instantaneous crank angle rad

θadvance Crank angle of advance rad

θspark,start Crank angle at which spark delivery is initiated rad

θspark,stop Crank angle at which spark delivery is terminated rad

ΔG Change in Gibbs free energy J/mol

ΔHj Enthalpy change for reaction j J/mol

References

[1] Nicoletti, G. (1995). The hydrogen option for energy: A review of technical, environmental
and economic aspects. International Journal of Hydrogen Energy, 20(10), 759-765.

[2] Veziroǧlu, T. N., & Şahin, S. (2008). 21st Century’s energy: Hydrogen energy system.
Energy Conversion and Management, 49(7), 1820–1831.

[3] Chaichan, M. T., & Abass, Q. A. (2010). Study of NOx Emissions of SI Engine Fueled with
Different Kinds of Hydrocarbon Fuels and Hydrogen, Al-Khwarizmi Engineering Journal, 6(2),
11-20
56

[4] Guzzella, L., & Onder, C. H. (2004). Introduction to Modeling and Control of Internal
Combustion Engine Systems, New York: Springer-Verlag Berlin Heidelberg.

[5] Maamri, R., Martshenko, A. P., Osetrov, O. O., Dubé, Y., Toubal, L., & Kodjo, A. (2013).
Analyze and Mathematical Modeling of the Combustion Process of One-Cylinder Spark Ignited
Hydrogen Fueled Engine, American Journal of Vehicle Design, V(2), 21–24.

[6] Rahman, M. M., Noor, M. M., Kadirgama, K., & Rejab, M. R. M. (2009). Study of air fuel
ratio on engine performance of direct injection hydrogen fueled engine. European Journal of
Scientific Research, 34(4), 506–513.

[7] Kuo, P. S. (1996). Cylinder Pressure in a Spark-Ignition Engine : A Computational Model.


Journal of Undergraduate Sciences, 3(Fall), 141–145.

[8] Blumberg, P. N., Lavoie, G. A., & Tabaczynski R. J. (1979). Phenomenological models for
reciprocating internal combustion engines, Progress in Energy and Combustion Science, 5(2),
123–167.

[9] Heywood, J. B. (1988). Internal Combustion Engines Fundamentals, Singapore: McGraw-


Hill.

[10] Ma, J., Su, Y., Zhou, Y., & Zhang, Z. (2003). Simulation and prediction on the performance
of a vehicle’s hydrogen engine. International Journal of Hydrogen Energy, (28), 77–83.

[11] Rahman, M. M., Hamada, K. I., Noor, M. M., Bakar R. A., Kadirgama, K. & Maleque, M.
A. (2010). In-Cylinder Heat Transfer Characteristics of Hydrogen Fueled Engine : A Steady
State Approach, American Journal of Environmental Sciences vol. 6 (2): 124-129.

[12] Mikulski, M., Wierzbicki S., & Piętak A. (2015). Zero-dimensional 2-phase combustion
model in a dual-fuel compression ignition engine fed with gaseous fuel and a divided diesel fuel
charge, Maintenance and Reliability, 17(1), 42–48.

[13] Vudumu, S. K. (2010). Experimental and computational investigations of hydrogen safety,


dispersion and combustion for transportation applications. Doctoral Dissertations. Paper 2171.

[14] Fagelson J. J., McLean W.J., & de Boer P.C.T. (1978). Performance and NOx emissions of
spark-ignited combustion engines using alternative fuels – quasi onedimensional modeling. I.
Hydrogen fueled engines. Combustion Science and Technology (18),47–57.

[15] Prabhu-Kumar, G.P., Nagalingam, B., & Gopalakrishnan, K.V.(1985). Theoretical studies
of a spark-ignited supercharged hydrogen engine, International Journal of Hydrogen Energy,
(10), 389 –397.

[16] Verhelst, S., & Sierens, R. (2007). A quasi-dimensional model for the power cycle of a
hydrogen-fuelled ICE, International Journal of Hydrogen Energy, (32), 3545–3554
57

[17] D’Errico, G. D., Onorati, A., & Ellgas, S. (2008). 1D thermo-fluid dynamic modelling of an
S.I. single-cylinder H 2 engine with cryogenic port injection. International Journal of Hydrogen
Energy, 33(20), 5829–5841.

[18] Knop, V., Benkenida, A., Jay, S., & Colin, O. (2008). Modelling of combustion and
nitrogen oxide formation in hydrogen-fuelled internal combustion engines within a 3D CFD
code. International Journal of Hydrogen Energy, 33(19), 5083–5097.

[19] Verhelst, S., & Sheppard, C. G. W. (2009). Multi-zone thermodynamic modelling of spark-
ignition engine combustion - An overview. Energy Conversion and Management, 50(5), 1326–
1335.

[20] Johnson N. Hydrogen as a zero-emission, high-efficiency fuel: uniqueness, experiments and


simulation (1997). 3rd International Conference on Internal Combustion Engines-97, Internal
combustion engines: experiments and modeling (Naples, Italy)).

[21] Adgulkar, D.D., Deshpande, N.V., Thombre, S.B., & Chopde, I.K. (2008). 3D CFD
simulations of hydrogen fuelled spark ignition engine. ASME Spring Engine Technology
Conference paper nr. ICES2008-1649 (Chicago, Illinois, USA).

[22] Al-Baghdadi M.A.R.S. (2004). Effect of compression ratio, equivalence ratio and engine
speed on the performance and emission characteristics of a spark ignition engine using hydrogen
as a fuel, Renewable Energy, 29, 2245–2260.

[23] Safari, H., Jazayeri, S. A., & Ebrahimi, R. (2009). Potentials of NOx emission reduction
methods in SI hydrogen engines : Simulation study. International Journal of Hydrogen Energy,
34(2), 1015–1025.

[24] Jones, W. P. & Lindstedt, R. P. (1988). Global Reaction Schemes For Hydrocarbon
Combustion, Combustion and Flame, 73, 233-249.

[25] Marinov, N. M., Curran, H. J., Pitz, W. J., & Westbrook, C. K. (1998). Chemical kinetic
modeling of hydrogen combustion under conditions found in internal combustion engines.
Energy & Fuels, 12(1), 78-82.

[26] Marinov, N. M., Westbrook, C. K., & Pitz, W. J. (1996). Detailed and Global Chemical
Kinetics Model for Hydrogen combustion. Transport Phenomena in Combustion, 1(March), 118-
129.

[27] Flagan R. C., & Seinfield J. H. (1988). Fundamentals of air-pollution control engineering
(Vol. 1), New Jersey:Prentice-Hall Inc.

[28] Karim, G. (2003). Hydrogen as a spark ignition engine fuel. International Journal of
Hydrogen Energy, 5, 569–577.

[29] White, F. (2010). Fluid Mechanics, New York :McGraw-Hill.


58

[30] Kumar P., Franchek M., Grigoriadis K., & Balakotaiah V. (2011). Fundamentals-Based
Low-Dimensional Combustion Modeling of Spark-Ignited Internal Combustion Engines, AIChE
Journal, September, 57(9),2472-2492.

[31] Bowman, C. T. (1975). Kinetics of pollutant formation and destruction in combustion.


Progress in Energy and Combustion Science, 1(1), 33–45.

[32] Khajuria P. R., Hydrogen-fuelled spark ignition engine: Its performance and exhaust
emission characteristics. Ph.D. Thesis, I.I.T. Delhi, India (1981).

[33] Kleinhenz, J., Sarmiento, C., & Marshall, W. (2012). Experimental Investigation of
Augmented Spark Ignition of a LO2 / LCH4 Reaction Control Engine at Altitude Conditions,
48th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, Atlanta, Georgia, 1–14.

[34] Woschni, G. (1967). A universally applicable equation for the instantaneous heat transfer
coefficient in the internal combustion engine, SAE Technical Paper 670931.

[35] Jilakara, S., Vaithianathan, J. V, Natarajan, S., Ramakrishnan, V. R., Subash, G., Abraham,
M., Unni, J.K., & Das, L.M. (2015). An Experimental Study of Turbocharged Hydrogen Fuelled
Internal Combustion Engine, SAE International Journal of Engines, 8(1), 314–325.

[36] Demuynck, J., De Paepe, M., Huisseune, H., Sierens, R., Vancoillie, J., & Verhelst, S.
(2011). On the applicability of empirical heat transfer models for hydrogen combustion engines.
International Journal of Hydrogen Energy, 36(1), 975–984. doi:10.1016/j.ijhydene.2010.10.059

[37] Wei, S. W., Kim, Y.Y., Kim, H.J., & Lee, J.T. (2001). A study on transient heat transfer
coefficient of in-cylinder gas in the hydrogen fueled engine, 6th Korea-Japan Joint Symposium
on Hydrogen Energy.

[38] Powell, R. W., Ho, C. Y., & Liley, P. E. (1966). Thermal conductivity of selected materials,
Report prepared at The Thermophysical Properties Research Center, Purdue University, Indiana,
1-168.

[39] Mohammadi, A., Shioji, M., Nakai, Y., Ishikura, W., & Tabo, E. (2007). Performance and
combustion characteristics of a direct injection SI hydrogen engine. International Journal of
Hydrogen Energy, 32(2), 296–304.

[40] Chan, K., Ordys, A., Volkov, K., & Duran, O. (2013). Comparison of Engine Simulation
Software for Development of Control System, Modelling and Simulation in Engineering (2013),
1-21.

[41] Das, L. M. (1991). Exhaust emission characterization of hydrogen-operated engine system:


Nature of pollutants and their control techniques. International Journal of Hydrogen Energy,
16(11), 765–775.
59

[42] Mathur, H., & Khajuria, P. (1984). Performance and emission characteristics of hydrogen
fueled spark ignition engine. International Journal of Hydrogen Energy, 9(8), 729–735.

[43] Verhelst, S., & Wallner, T. (2009). Hydrogen-fueled internal combustion engines. Progress
in Energy and Combustion Science, 35(6), 490–527.

[44] Subramanian, V., Mallikarjuna, J. M., & Ramesh, A. (2007). Intake charge dilution effects
on control of nitric oxide emission in a hydrogen fueled SI engine, International Journal of
Hydrogen Energy (32), 2043–2056.

[45] Jabbr, A. I., Vaz, W. S., Khairallah, H. A., & Koylu, U. O. (2016). Multi-objective
optimization of operating parameters for hydrogen-fueled spark-ignition engines. International
Journal of Hydrogen Energy, 41(40), 18291–18299. doi:10.1016/j.ijhydene.2016.08.016

[46] White, C. M., Steeper, R. R., & Lutz, A. E. (2006). The hydrogen-fueled internal
combustion engine: a technical review. International Journal of Hydrogen Energy, 31(10),
1292–1305.

[47] Wallace, J. S., & Ward, C. A. (1983). Hydrogen as a fuel. International Journal of
Hydrogen Energy, 8(4), 255–268.

View publication stats

You might also like