ignition delay study of next generation alternative jet

34
IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET FUELS IN A RAPID COMPRESSION MACHINE BY KYUNGWOOK MIN THESIS Submitted in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering in the Graduate College of the University of Illinois at Urbana-Champaign, 2015 Urbana, Illinois Adviser: Associate Professor Tonghun Lee

Upload: others

Post on 21-Jan-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET FUELS IN A RAPID

COMPRESSION MACHINE

BY

KYUNGWOOK MIN

THESIS

Submitted in partial fulfillment of the requirements

for the degree of Master of Science in Mechanical Engineering

in the Graduate College of the

University of Illinois at Urbana-Champaign, 2015

Urbana, Illinois

Adviser:

Associate Professor Tonghun Lee

Page 2: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

ii

ABSTRACT

Alternative fuels have been widely and actively investigated recently to alleviate an impending

energy crisis. Rising environmental, economical, and political concerns requires employments of

alternative sources of energy other than conventional fossil fuels. Alternative aviation fuels from

diverse bio-feedstock have been introduced to reduce dependency on fossil fuels and

environmental effects. However, combustion characteristics and properties of the newly

developed fuels are not yet comprehensively understood. Recent studies have been examined to

obtain combustion characteristics of alternative aviation fuels, as well as physical and chemical

properties. The primary goal of this study is to determine the ignition delay time of jet fuels of

interest; conventional, alternative, and surrogate blends. By measuring the pressure trace of

autoignition in rapid compression machine, ignition delay time is captured through the pressure

derivative. Category A fuels represent conventional jet fuels. Three of the category A fuels of

interest are Jet A(Jet A-2), the nominal commercial aviation fuel, JP-8(Jet A-1) and JP-5(Jet A-3),

both of which are conventional military jet fuels. Fuels in category C are the surrogate fuels with

specific targeted physical properties, or chemical composition. They are either produced from

bio feedstock, or blend of them with conventional fuels or surrogate components. Amyris

Farnesame, Gevo ATJ(C-1), blend of tetradecane and trimethylbenzene (C-2), blend of JP-5 and

farnesane (C-3), blend of Sasol IPK and Gevo ATJ (C-4), blend of decane and trimethylbenzene

(C-5) are tested through the study. A rapid compression machine at the University of Illinois at

Urbana-Champaign is used for testing, with the direct test chamber charge preparation method.

DTC configuration has advantages in avoiding thermal decomposition and controllability.

Page 3: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

iii

ACKNOWLEDGMENTS

I would like to express sincere appreciation to my great adviser, Professor Tonghun Lee for his

guidance and assistance. I am also grateful to all of my lab mates, sincere friends, especially to

Daniel Valco for sharing his expertise and knowledge. Finally, I thank to my family who always

have been supportive for me.

Page 4: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

iv

Table of Contents

1. INTRODUCTION .................................................................................................................................... 1

1.1 Hydrocarbons ...................................................................................................................................... 1

1.2 Aviation fuels ...................................................................................................................................... 1

1.3 Alternative Fuels ................................................................................................................................. 2

1.4 Fuel surrogate...................................................................................................................................... 3

1.5 Ignition delay ...................................................................................................................................... 3

2. EXPERIMENTAL METHOD .................................................................................................................. 5

2.1 Rapid Compression Machine .............................................................................................................. 5

2.2 Testing fuels ...................................................................................................................................... 12

3. RESULTS AND DISCUSSIONS ........................................................................................................... 18

3.1 Ignition delay time of category A fuels ............................................................................................ 18

3.2 Ignition delay time of category Farnesane and Jet C-1 ..................................................................... 20

3.3 Ignition delay time of Jet C-2 and Jet C-5 ........................................................................................ 22

3.4 Ignition delay time of Jet C-3 ........................................................................................................... 23

3.5 Ignition delay time of Jet C-4 ........................................................................................................... 24

4. CONCLUSION ....................................................................................................................................... 26

4.1 Recommended Future study.............................................................................................................. 26

References ................................................................................................................................................... 28

Page 5: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

1

1. INTRODUCTION

1.1 Hydrocarbons

Conventional fuels are mainly fossil fuels, composed of hydrocarbon components. Energy

stored in carbon bonds release bond energy when they are oxidized, or decomposed to smaller

product species. Accordingly, number of carbons, bond order and their structure affect ignition

characteristics, as well as physical, chemical properties of the fuels. If the carbon numbers and

functional groups are the same, species those who have lower order have less bond energy which

means easier to initiate combustion reaction. Heavier species or longer chain molecules have

higher boiling point due to low volatility; however, these species are usually more reactive. In

addition, it is known that heavily branched species are less reactive than normal, or lightly

branched species [1]. Regarding chain length, it is observed that the longer chains tend to be

more reactive [2]. Through gas chromatography-mass spectroscopy (GCMS), chemical species

contained in the fuels can be investigated, reaction characteristics and combustion behavior can

then be estimated.

1.2 Aviation fuels

Most of the fuels we use today come from the distillation of crude oil. Hence, distilled fuels

are a mixture of diverse hydrocarbon species. Three representative transportation fuels are

gasoline, jet fuel and diesel. Fuels are obtained from fractional distillation, by their different

volatility due to diverse molecular weight. Gasoline is lightest aviation fuel among three,

consists of light components, typically C6 to C12. Alkanes and aromatics account for most of the

Page 6: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

2

species. For commercial use, several additive may added for specific purposes. Gasoline is main

fuel of spark ignitions engines. For spark ignition engines, autoignition is referred to knocking,

which is detrimental for the engine system. Octane number is indicator for fuels propensity of

autoignition, or the reactivity. Octane number of the fuels is numerically generated by comparing

autoignition character fuels with two reference species; iso-octane(2,2,4 trimethylpentane) and n-

heptane. Higher the octane number is, the less likely to knock, in other words, less reactive.

Diesel fuels are heavy transportation fuels, blends of C9 to C20. Average molecular weight is

heavier than gasoline, as well as energy density. Cetane number is index referring self-ignition

propensity which is similar to octane number, however, the higher cetane number fuels have

higher reactivity. Cetane number is obtained experimentally by comparing self-ignition behavior

of the fuels with blends of two reference fuel components; n-cetane(hexadecane) and

heptamethylnonane(HMN). Aviation fuels, or jet fuels, are the transportation fuels that are in

between gasoline and diesel, called kerosene. Molecular weight, chain length, boiling point,

energy density and other properties are in between gasoline and diesel, both octane number or

cetane number are used to refer their reactivity. For aviation application, specific additives are

added to satisfy commercial or military standard of fuel properties, such as flash point, freezing

point. [3, 4]

1.3 Alternative Fuels

By far, fossil fuels are the most common type of energy source, especially for transportation

fuels. Energy density of fossil fuels is higher than other types of sources in terms of volume and

mass, due to chemical energy stored in organic species of fossil fuels. High energy density

results in longer driving range, which is one of the most important factors of vehicle

characteristics. However, constantly increasing consumption of fossil fuel energy sources will

Page 7: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

3

inevitably result in depletion of fossil fuel feedstock. Additionally, extraction of fossil fuels from

underground sources increases the overall number of carbon in the ecosystem, in other words,

increasing carbon footprints, which accounts for global warming. The term ‘alternative’ can be

used widely, for referring any energy sources other than conventional fossil fuels. For this study,

however, alternative fuel means the hydrocarbon based fuels produced from alternative sources,

in short, non-petroleum. Coal, natural gas, oil shale and biomass are the potential alternative

source of liquid fuels.

1.4 Fuel surrogate

Fuel Surrogates are basically blends of several organic species. Specific species are chosen

to emulate physical and/or chemical properties, to emulate the combustion characteristics of

parent fuels. Specific characteristics of the surrogate that aims to match a parent fuel is called a

target. Fuel surrogates are made to precisely examine the mechanism, simulations, or in-lab tests.

Simulation results and test measurements are compared to each other in order to examine validity

of the surrogates. Several surrogates for jet fuel have been evaluated through experiments and

simulations [5-9]. Farnesane is a sole component fuel, but its DCN is close to of S-8, synthetic

JP-8 aviation fuel, so it has been studied as a surrogate for the JP-8 [10]. Category C fuels are the

surrogates made to examine specific cases of fuel properties, which will be discussed further in

§2.1.2.

1.5 Ignition delay

Ignition delay time is the time interval between the moment when system is at autoignition

condition and the onset of combustion reaction [11]. Ignition delay time consists of both physical

and chemical delay. Vaporization of injected fuels, heat transfer, mixing and other mechanics

accounts for physical delay, and the required time for the fuel-air mixture to start chemical

Page 8: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

4

reactions accounts for the chemical delay [12, 13]. However, for this study, homogeneous

premixed charge is assumed through DTC preparation configuration, so the chemical delay is the

main interest of measurement. Ignition delay time has inverse relation with reactivity of the fuels.

Chemical reaction of combustion is attributed to the formation of free radicals, which is more

likely to happen with reactive species [14]. The time measurement starts when the premixed

charge is fully compressed and ends with a combustion event. Defining onset of combustion

event is difficult. Light emission, change in heat release rate, temperature or pressure rise can be

the marker of combustion reaction [12]. Among these, using pressure measurement is known to

be most reliable and straightforward. For this study, maximum pressure raise rate point is set to

be the onset of combustion.

Page 9: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

5

2. EXPERIMENTAL METHOD

2.1 Rapid Compression Machine

Ignition delay times of the fuels of interest have been measured using a Rapid Compression

Machine (RCM) at the University of Illinois at Urbana-Champaign. The RCM is a

pneumatically-driven and hydraulically-stopped machine which emulates a single stroke of

engine devices [15]. A RCM is used to measure ignition delay time of fuels of interests, by

observing pressure trace of autoignition behavior. Previous studies of the device and specific

experimental setup are discussed in this section.

2.1.1 Previous studies

RCM is widely used for combustion studies, especially for auto ignition phenomena. Early

20 centuries, apparatus that compresses charged gas using momentum of weight have been

introduced[16]. For modern RCM, pneumatically driven piston is widely used [17-21], however,

stopping mechanisms vary. Donovan et al.[22] designed a RCM with a free sabot piston, which

stops when its coned shape piston fits to the end of stroke range. The RCM at University of

Science and Technology at Lille has cam structure for stopping [23]. The RCM used in this study

uses hydraulic pistons and a control valve, which is the most widely used[15, 24, 25].

Most of the RCMs are used for investigating combustion kinetics of fuels. Regardless of the

driving and breaking mechanism, compression stroke is the process of interest. Focusing on

chemical kinetics, zero-dimensional modeling should be valid when combustion occurs. Several

studies have assessed the validity of this approach. The most important assumption is an

Page 10: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

6

adiabatic core assumption, which assumes uniform temperature after the compression. Mittal and

Sung [26] assessed the aerodynamics of the compression stroke using CFD modeling and PLIF

imaging of acetone. Their investigation suggests the need to use a creviced piston for a

homogeneous reaction core, which reducesthe roll-up vortex at the corners during compression.

Griffiths et al. [27] also investigated temperature fields of RCM after the compression stroke,

concluding spatial variation of temperature is trivial, and there exists an adiabatic gas core.

Other than adiabatic reaction core, heat loss of the reaction region has been investigated [28,

29]. Considering the compression stroke of the modern RCM takes less than a hundred

milliseconds, and the ignition delay of the fuels usually less than hundreds of millisecond, heat

losses during the time affect the condition of reaction core region. Wurmel et al. [28] showed

that the diluent gas composition affects ignition delay time as well as the adiabatic core

temperature, especially helium, which result in greater heat loss than other diluent gases.

2.1.2 Device specifications

The RCM at University of Illinois, Urbana-Champaign, which is used in in this study, has a

hydraulic stopping mechanism. Figure 1shows a schematic configuration of RCM in University

of Illinois, Urbana Champaign. An air reservoir tank is charged to 140 bar from an external air

compressor when running a test. The external compressor is a C-Aire 5HP, 80 gallon two stage

air compressor (model number : A050V080-1230). The hydraulic chamber is connected to an oil

pump (Star Hydraulics, CP13-250-2000) and filled with heavy, viscous mineral oil to hold the

piston before it is ‘fired.’ Solenoid valve (Parker, 71216SN2BL00N0C111P3) is installed

between the oil pump and the hydraulic chamber, and is controlled by National Instrument

Labview. The valve is closed when pneumatic chamber is charging, and once it is charged to

140 bar and the combustion chamber is filled with the test mixture charge, the operator opens the

Page 11: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

7

solenoid valve to drain the mineral oil in hydraulic chamber back into pump. Since mineral oil is

not holding the piston anymore, the whole piston rod moves forward (toward combustion

chamber) and the charged mixture in combustion chamber is compressed by the piston in

combustion cylinder.

The combustion chamber and cylinder is heated with band heaters for fuel vaporization. For

this study, temperature of the chamber and cylinder is retained at 125℃; however, temperature

slightly varies due to compression, ignition and evacuation. Six thermocouples (Omega,

TMQSS-125U-4) are attached to the cylinder and chamber wall to control the band heater power

through a Labview PID controller. Chamber and cylinder are insulated with a custom made glass

fiber insulation jacket and the combustion cylinder and chamber are covered with safety shield

made of bulletproof acryl as shown in Figure 2.

Figure 1 Schematic diagram of RCM

Air Reservoir

Pneumatic

Cylinder

Hydraulic

Chamber Combustion

Chamber

Combustion

Cylinder

Oil Pump

Oxyg

en

Dry

Air

Page 12: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

8

Figure 2 RCM at University of Illinois, Urbana Champaign

2.1.3 Direct test chamber

Direct test chamber (DTC) configuration is used for the entire study. For conventional

premixed charge test, a mixing vessel is required. Diluent and oxidation gas are mixed with fuels

in the mixing vessel and fed into the test chamber. To ensure homogeneity of the mixture, the

vessel is usually heated for evaporation and mixing mechanism such as magnetic stir is equipped.

However, especially when dealing with liquid fuels with low volatility, the vessel has to be

heated to somewhat high temperature which can result in thermal decomposition, or pyrolysis, of

the fuel component. In addition, to fill the test chamber with certain pressure, pressure of the

mixing chamber should be higher than test chamber pressure, heating temperature requirement

Page 13: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

9

raises. Temperature and pressure drops in feeding lines are additional problems that need to be

resolved when using an external mixing chamber.

Allen et al. suggested the DTC configuration which can effectively solve the limitation of

using a mixing vessel [30]. A fuel injector is directly installed into the test chamber section of the

RCM, diluent and oxidation gas are charged into the chamber separately. A schematic

configuration and picture of DTC set up is shown in Figure 3. After the combustion chamber and

cylinder are heated to proper temperature for fuels to fully evaporate, diluent and oxidant gas are

charged. For this study, synthetic dry air composed of 79% nitrogen and 21% oxygen is used for

all fuels. Filling pressure determines the compressed pressure, together with the compression

ratio of the piston stroke, which can be adjusted by adding additional shims between the

hydraulic cylinder and combustion cylinder, or hydraulic cylinder and pneumatic cylinder. For

Band Heater

Combustion Chamber

Hydraulic Cylinder

Fuel Injector

Piston

Optical Window

Fuel Accumulator

Figure 3 DTC configuration

Page 14: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

10

this study, compression ration varies from the lowest 4.52 to the highest 7.17. Pc has been

retained to 20 bar throughout this study. Initial charged pressure and final compressed pressure

determine the compressed temperature, using isentropic relation as following:

∫𝛾

𝛾 − 1

𝑇𝑐

𝑇0

𝑑𝑇

𝑇= 𝑙𝑛

𝑃𝑐𝑃0

where T and P is Temperature and Pressure, c denotes compressed condition, 0 denotes initial

condition, and γ is temperature dependent specific heat ratio [24]. Specific heat ratio for diluent,

oxidant and fuel are calculated from NASA polynomials form coefficients of each gases and

fuels, from NIST webbook and Thermochemical properties of jet fuels, version 5 report [31].

Once the diluent and oxidant are charged into the chamber and cylinder, fuel is injected to the

chamber with a fuel injector. The fuel accumulator is pressurized to 30 bar, feeding the fuels to a

Bosch gasoline direct injection fuel injector (Bosch, 13537591623). For different fuels, different

injectors are installed after calibration process for precise amount of injection of fuels, between

15~100 mg, depending on types of fuels and stoichiometric ratio. For precise control, fuel is

injected with pulses less than 1 millisecond, using a custom made injector controller box, which

sends a DC signal to the fuel injector. Stoichiometric ratios of unity, 0.5 and 0.25 have tested

throughout the study; however, for fuels with low reactivity, for instance Gevo ATJ, no ignition

occured at ϕ =0.25. To ensure homogeneous evaporation and mixing, 4 minutes of preparation

time is required after the fuel is injected [30]. Once the charge is assumed to be uniformly mixed,

the piston is fired and mixture is compressed to reach autoignition conditions. Pressure trace

history of the autoignition behavior is monitored and logged using an incoming signal from a

Kistler pressure transducer (Kistler 6125B21), amplified through dual mode signal amplifier

(Kistler, type 5010).

Page 15: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

11

2.1.4 Post processing of the data

Obtained signal of pressure trace have been filtered using a low-pass Butterworth filter, with

5000 Hz cutoff frequency. Top dead center, TDC, is defined to be the same time as the moment

at the end of compression. For most of the test cases with distinguishable ignition delay times,

this moment can be found easily, where time derivative of the pressure trace become zero for the

first time. Figure 4 shows a typical pressure trace and time derivative, with TDC identified in

each case. However, when chemical reactions occur earlier during compression, early knock or

early ignition is observed, and the pressure derivative does not reach to zero. In this case, the

moment when pressure derivative reaches local minimum has taken to be the TDC, though, it

might not be the exact TDC. Figure 5 is exemplary case if this case, early ignition is observed

during the compression stroke.

Ignition delay time in this study is defined to be the time delay between TDC and maximum

pressure rise moment. The moment when the pressure derivative has local maxima is chosen to

be an instance of ignition. Multistage ignition phenomena is observed for some of the cases, if so,

TDC

TDC

Figure 4 Pressure trace(left) and time derivative of pressure trace(right) of Jet A-1, Tc=665K, Pc=20 bar, phi=1.0

Page 16: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

12

TDC

TDC

τ1 τ2

τ

TDC

ignition delay time can be separated to each stage; first and second stage ignition time delay,

those combined will give the overall ignition delay time. Figure 6 is one of the exemplary

conditions where multistage ignition behavior is prominently observed. For a few cases, third

stage ignition is observed; however, it is hardly distinguishable so will not be regarded as a

significant stage. Unless otherwise stated, ignition delay time refers overall ignition delay time

throughout this study.

Figure 5 Pressure trace(left) and time derivative of pressure trace(right) of Jet A-1, Tc=709K, Pc=20 bar, phi=0.5

Figure 6 Exemplary pressure trace of multistage ignition, Farnesane, Tc=646K, Pc=20 bar, phi=0.5

Page 17: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

13

2.2 Testing fuels

Conventional, alternative and surrogate aviation fuels have been chosen for the study. Jet

fuels that are currently used widely are classified as a category A fuels, of which combustion

properties as well as physical, chemical properties are fairly known. Category C fuels are

alternative candidates or surrogates components of jet fuels, which are classified to be

investigated for their validity of targeted combustion characteristic. Alternative or synthetic fuels

that have been examined for validity are category B fuels [32]. Three representative category A

fuels and some category C fuels are the objects of interest in this study.

2.2.1 Category A fuels

Jet A-1, Jet A-2, Jet A-3 are the most widely used aviation fuels. Nominal case of JP-8 and

extreme case of ‘best’ and ‘worst’, whose properties are on the edge of requirements. The best

case fuel, Jet A-1(POSF 10264), and Jet A-3(POSF 10289), ‘worst case’ fuel has converse

properties. Jet A-2(POSF 10325)is nominal case fuel and has the properties in between the two

aforementioned fuels [33]. Properties of the category A fuels are on Table 1. Normalized GCMS

data showing the chemical composition of the fuels are show in Figure 7. The three category A

fuels have few difference in composition; however, Jet A-3 shows a slightly narrow boiling

range, which classifies it as a narrow cut fuel. All fuels contain iso-paraffins, n-paraffins, cyclo-

paraffins, and aromatics. Contents of the aromatics are sometimes restricted due to emission

issues.

Recent studies about JP-8(Jet A-1), Jet A(Jet A-2), JP-5(Jet A-3) and their surrogates have

been done for ignition delay time measurement in RCM[15, 30, 34, 35], shock tube[36-38] and

calculation through kinetic modeling and simulations[39-42].

Page 18: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

14

Jet A-1

Jet A-3

Jet A-2

n-C8

n-C9

n-C10 n-C11

n-C12 n-C13

n-C14

n-C15

n-C16

Table 1: Properties of Category A fuels

Fuel POSF Cetane MW H/C ratio iso-

paraffins n-paraffins

cyclo-

paraffins aromatics

Index (g/mole) (weight %)

Jet A-11 10264 ~49.6 141.4 2.011 39.69 26.82 20.08 13.41

Jet A-21 10325 ~50 147.6 1.961 29.45 20.03 31.86 18.66

Jet A-31 10289 ~40.9 156.1 1.868 18.14 13.89 47.38 20.59

1 Category A fuels data are established by Hai Wang, Stanford

Figure 7 GCMS data of Category A fuels

Page 19: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

15

2.2.2 Category C fuels and Farnesane

Farnesane and Gevo ATJ (Jet C-1) are the alternative aviation fuels from bio-feedstock.

Farnesane is produced from biomass sugars, so is called a direct sugar to hydrocarbon (DSHC)

fuel, and its sole component is 2,6,10 trymethyl dodecane, or iso-pentadecane(C15). Farnesane

has a hydrogen to carbon molar ratio of 2.13, which is slightly higher, but close to conventional

category A fuels, at around 2 or below. The higher DCN of Farnesane is expected to result in

Jet C-1

TMB

i,n-C10

i-C12

i-C14

i-C16

Jet C-2

Jet C-3

Jet C-4

Jet C-5

i-C15

i-C12 i-C16

TMB

Figure 8 GCMS data of Category C fuels

Page 20: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

16

shorter ignition delay time than category A fuels. Won et al.[10] investigated that Farnesane and

found is about the same DCN as S-8 (POSF 4734) synthetic aviation fuel. Dooley et al.[42]

investigated that at high temperatures, above 1000K, S-8 and JP-8 have similar ignition

characteristics; however, at low temperature regions, shorter ignition delay time of S-8 has been

shown by kinetic modeling and shock tube experimental data. 20%/80% of Amyris Farnesane/Jet

A-1 blend has been tested on PW615F jet engine without any problem in operability and is

comparable to Jet A-1 [43].

Jet C-1 fuel, or Gevo ATJ, is derived from an alcohol, butanol, which is obtained through the

fermentation process of cellulosic biomass feedstock. It has an irregular composition compared

to conventional jet fuels. Two heavily branched paraffinic species mostly consist the fuel; 78.3%

iso-dodecane(2,2,4,4,6 or 2,2,4,4,6 pentamethyl heptane) and 16.3% iso-cetane(2,2,4,4,6,8,8 or

2,2,4,4,6,6,8) by mass. These heavily branched paraffinic species account for extremely low

DCN(~15) of Gevo. Zhu et al. [44] measured ignition delay times in a shock tube and the results

show a longer delay time of Gevo compared to JP-8, at high temperatures above 1000K.

According to the manufacturing company Gevo, several engine and aviation tests have been

done with Gevo and JP-8 blends without noticeable difference in performance [45].

Jet C-2 fuel is ‘bimodal’ mixture of tetradecane(C14) and 1,3,5-trimethylbenzene(TMB). The

blend is made to examine the impact of particular fuel chemistry of volatile species, TMB. Jet C-

3 is blend of Farnesane and JP-5, which has a high viscosity. Jet C-4 is a blend of Gevo ATJ and

Sasol IPK fuels, which has a wide boiling range and low DCN. On the other hand, Jet C-5 has

narrow boiling range, as shown in Figure 8. Properties of these fuels are listed in Table 2.

Page 21: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

17

Table 2 Properties of category C fuels

Fuel posf # Cetane MW H/C ratio iso-paraffins n-paraffins cyclo-paraffins aromatics

Index (g/mole)

Jet C-1 11498 ~15 178.5 2.16 99.62 <0.01 0.05 <0.01

Jet C-2 12223 N/A 181.9 2.08 77.51 5.16 0.07 17.05

Jet C-3 12341 N/A 179.6 1.98 45.19 9.17 31.72 13.61

Jet C-4 12344 N/A 162.2 2.18 98.94 0.23 0.43 0.39

Jet C-5 12345 N/A 135.4 1.93 51.58 17.66 0.07 30.68

Amyris Farnesane 10370 ~59.1 212.4 2.13 100 - - -

(weight %)

Page 22: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

18

3. RESULTS AND DISCUSSIONS

3.1 Ignition delay time of category A fuels

Ignition delay times measured in the RCM are evaluated through averaging a minimum of

three repetitive measurements at the same condition. Measurements have been done from the

lowest temperature that the device can reach, to the highest temperature where identification of

TDC is possibly, before early knock is observed. Compressed pressure (Pc) has been controlled

to be at 20 bar by altering the initial pressure (P0) and compression ratio. Overall results for the

category A fuels are in Figure 9 below.

Figure 9 Category A fuels ignition delay time

Page 23: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

19

3.1.1 Fuel Comparison

As shown in Figure 7, chemical species in category A fuels of interest are mostly the same. It

is shown that there are small discrepancies among the three selected fuels, especially at lean

conditions. However, measured ignition delay times of the three fuels are similar considering the

error range of the measurements. The results correspond to a previous study done by Kumar et al.

[46] which states that JP-8(Jet A-1) and Jet A(JetA-2) have similar auto ignition behavior.

Figure 10 Category A fuels ignition delay time [2]

Page 24: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

20

3.1.2 Temperature and stoichiometric ratio dependency

Ignition delay time is measured to be shorter at higher temperature as shown in Figure 9 and

Figure 10. A flat slope in the Arrhenius plot is observed, which means overall ignition delay time

exponentially decreases with increasing temperature. Prominent NTC behavior is hard to observe,

except for the ϕ=0.5 condition where slope decreases with increasing temperature at over 700K.

The temperature range of the testing conditions cover the NTC regime on Kumar and Sung [46],

however, it is hard to see the NTC behavior, because the compressed pressure is higher in this

study. Overall ignition delay times are longer with leaner conditions at the same temperature,

which corresponds to previous studies that examined these kinds of the fuels.

3.2 Ignition delay time of category Farnesane and Jet C-1

Farenesane and Jet C-1(Gevo ATJ) are the neat fuels, whereas C-2 throughC-5 fuels are

blended fuels. Comparison of ignition delay times for these two fuels and Jet A-2 are in Figure

11. It is possible to get only few data points for the fuels, since Jet C-1 does not ignite at ϕ=0.25

lean condition and early knocking which make it difficult to determine TDC has observed on

Farnesane. Figure 11 has the measured overall ignition delay time for Jet C-1 and Farnesane,

with Jet A-2 for comparison. Ignition delay times of Farnesane are shorter than Jet A-2, whereas

of Jet C-1 are significantly longer than Jet A-2. This result corresponds to the DCN of the fuels,

which refers to the reactivity. Inverse correlation of DCN and ignition delay time has been

observed from several previous studies [10, 46, 47]. Comparing Jet A-2 and Farnesane, the

absence of aromatics and the longer chain of the Farnesane accounts for greater reactivity. At

low temperature conditions, where the tests have been done, oxidation reaction of aromatics is

slower than open chain species due to olefin formation and isomerization pathway. The long

carbon number chain of 15 and relatively light degree of branching of Farnesane also can explain

Page 25: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

21

its greater reactivity [33]. Jet C-1 has a considerably low DCN of 15, which results in longest

ignition delay times among the fuels of interest examined in this study. The two main species of

Jet C-1- pentamethly heptane and heptamethyl nonane- are heavily branched species, which have

low reactivity at low temperature conditions. It is also found that Jet C-1 enters the NTC regime

earlier at lower temperature than other fuels, especially at lean condition.

Figure 11 Ignition delay time of Jet A-2, Farnesane, Jet C-1

Page 26: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

22

3.3 Ignition delay time of Jet C-2 and Jet C-5

Figure 12 Ignition delay time of Jet C-2, Jet C-5 and Jet A-3

Jet C-2 and Jet C-5 are both mixtures of two main components; iso-tetradecane and

trimethyl benzene and iso-decane and trimethyl benzene, respectively. These two blends were

chosen to represent two extreme case of boiling range. Jet C-2 has an extremely asymmetric

boiling range, on the other hand, Jet C-5 has flat boiling range, in other words, all the fuel boils

at same temperature. These aspects can be seen on Figure 8 where peaks of Jet C-2 are separated

apart from each other, whereas peaks of C-5 are in close range. Jet A-3 is chosen to be the

reference of conventional jet fuel to compare with Jet C-2 to C-5. The average molecular size of

Jet C-2 (C12.9 H27.2) is greater than of Jet A-3 (C11.2 H12.0), whereas Jet C-5 has the smallest

Page 27: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

23

average molecules (C9.7 H19.7). Fuels with larger molecules tend to be more reactive due to longer

chain, because there is more chance of components oxidation. Comparing these three fuels,

aromatic content is a more important factor of reactivity. Aromatic compounds are known to

have low reactivity especially at low temperature condition. Jet C-2 has the least amount of

aromatics as shown in Figure 8, followed by Jet A-3 and Jet C-5 in order. High aromatic content

in Jet C-5 (27 volume %) accounts for its low reactivity and long ignition delay time as shown in

Figure 12. However, discrepancies become small under high temperatures and lean conditions.

3.4 Ignition delay time of Jet C-3

Figure 13 Ignition delay time of Jet C-3, Farenesane and Jet A-3

Page 28: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

24

Jet C-3 is blend of Farnesane (36 vol %) and Jet A-3 (64 vol %). Ignition delay time and

reactivity have been predicted to be in between Farnesane and Jet A-3, and the measurement

results match with the expectation as shown in Figure 13. The results show small discrepancies

among the fuels at high temperature, lean condition. The relatively high reactivity of Farnesane

and Jet C-3 deter testing at higher temperature because of early knock, which makes it difficult

to determine the moment of TDCor the end of compression.

3.5 Ignition delay time of Jet C-4

Jet C-4 is a blend of Jet C-1(Gevo ATJ) and Sasol IPK(Iso-Paraffinic Kerosene), at a ratio of

60:40 vol %, repectively. Since Jet C-1 has an extremely low DCN, Jet C-4 also is expected to

Figure 14 Ignition delay time of Jet C-4, Jet C-1 and Jet A-3

Page 29: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

25

have low reactivity. Sasol IPK has low contents of aromatic species; however, has a wide range

of species as shown in Figure 8. Jet C-4 is blended to have a low DCN and conventional boiling

range. Measured ignition delay times of Jet C-4 are compared with Jet C-1 and Jet A-3 in Figure

14. At unity equivalence ratio condition, Jet C-4 acts more like Jet A-3 rather than Jet C-1,

especially at higher temperatures. However, at leaner conditions, Jet C-4 shows longer ignition

delay time than Jet A-3, especially at higher temperature. The relatively longer ignition delay

time of Jet C-4 compared to Jet A-3 is prominent at leaner mixtures. Ignition delay times are not

as long as Jet C-1, which has extremely low DCN, and are still significantly longer than

conventional category A fuels due to heavily branched iso-dodecane and iso-cetane in Jet C-4.

Page 30: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

26

4. CONCLUSION

Three representative conventional jet fuels and six alternative fuels and surrogates have

been tested in rapid compress machine to measure ignition delay time. Jet A-1, Jet A-2, and Jet

A-3 fuels shows similar ignition delay times across the temperature and equivalence ratios

examined, with small discrepancies due to difference in chemical species group. Alternative

fuels and surrogate blends with irregular compositions show their own unique combustion

characteristics. Fuels with longer and highly branched components, with less aromatics, and

heavier average molecular weight tend to be more reactive. Blends which have two components

fuel show combustion characteristics is in between the two neat components.

4.1 Recommended Future study

All tests in this study have been done at a compressed pressure of 20 bar and three different

equivalence ratios, but was limited to low temperature conditions. Lowering the compressed

pressure may help understand pressure dependency. Also, at lower pressure, higher temperature

condition tests could be run where early knock has observed at 20bar. However, the RCM has

temperature limits due to its mechanical limitation. Intermediate and high temperature conditions

above 1000K can be reached with a shock tube, which is currently undergoing construction now

at the University of Illinois at Urbana-Champaign. Once the device is built, further

measurements at higher temperature condition can be conducted.

To understand the influence of chemical composition groups, a study of single components

and their blends would be helpful. Components with different degrees of branching at a similar

molecular weight might be tested and compared to inspect effects of branching in chemical

Page 31: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

27

studies. Previous studies of the components can be also compared directly or after scaling for the

test conditions.

This studies primary focus was on the building the RCM, its setup, and then conducting

preliminary measurements. Deeper studies and analysis on chemical reactions, mechanisms will

be done to comprehensively understand combustion behavior of fuels of interest. Some of the

fuels of interest in this study have relatively simple components, whose test results can be

validated through comparison with mechanism simulation results.

Page 32: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

28

References

1. Francis A. Carey, R., M. Giuliano, Organic Chemistry. 2011(Eigth edition).

2. Project, A.P.I.R., Knocking Characteristics of Pure Hydrocarbons. ASTM Special Technical

Publication No.225, 1916.

3. Pulkrabek, W.W., Engineering Fundamentals of the Internal Combustion Engine. 2004(Second

Edition).

4. Turns, S.R., An Introduction to Combustion: Concepts and Applications. 2000: McGraw-Hill.

5. Colket, M., et al. Development of an experimental database and kinetic models for surrogate jet

fuels. in 45th AIAA Aerospace Sciences Meeting and Exhibit. 2007.

6. Violi, A., et al., Experimental formulation and kinetic model for JP-8 surrogate mixtures.

Combustion Science and Technology, 2002. 174(11-2): p. 399-417.

7. Dooley, S., et al., A jet fuel surrogate formulated by real fuel properties. Combustion and Flame,

2010. 157(12): p. 2333-2339.

8. Honnet, S., et al., A surrogate fuel for kerosene. Proceedings of the Combustion Institute, 2009.

32 I: p. 485-492.

9. Dooley, S., et al., The combustion kinetics of a synthetic paraffinic aviation fuel and

fundamentally forumlated, experimentally validated surrogate fuel. Combustion and Flame, 2012.

159: p. 3014-3020.

10. Won, S.H., et al., The combustion properties of 2,6,10-trimethyl dodecane and a chemical

functional group analysis. Combustion and Flame, 2014. 161(3): p. 826-834.

11. Allen, C.M., et al., Ignition Characteristics of Diesel and Canola Biodiesel Sprays in the Low

Temperature Combustion Regime. Energy & Fuels, 2011. 25(7): p. 2896-2908.

12. Assanis, D.N., et al., A Predictive Ignition Delay Correlation Under Steady-State and Transient

Operation of a Direct Injection Diesel Engine. Journal of Engineering for Gas Turbines and

Power, 2003. 125(2): p. 450-457.

13. Spadaccini, L.J. and J.A. Tevelde, Autoignition characteristics of aircraft-type fuels. Combustion

and Flame, 1982. 46: p. 283-300.

14. Pilling, M.J., Low-temperature combustion and autoignition. 1997: Elsevier New York.

15. Allen, C.M., Advanced Rapid Compression Machine Test Methods and Surrogate Fuel Modeling

for Bio-Derived Jet and Diesel Fuel Autoignition, in Mechanical Engineering. 2012, Michigan:

East Lansing.

16. Falk, K.G., The ignition temperatures of gaseous mixtures. Second paper. Journal of the

American Chemical Society, 1907. 29(2): p. 1536-1557.

17. Affleck, W.S. and A. Thomas, An opposed piston rapid compression machine for preflame

reaction studies. Proceedings of the Institution of Mechanical Engineers, 1969. 183(1): p. 365-

385.

18. Beeley, P., J.F. Griffiths, and P. Gray, Rapid Compression Studies on Spontaneous Ignition of

Isopropyl Nitrate Part II: Rapid Sampling, Intermediate Stages and Reaction Mechanisms.

Combustion and Flame, 1980. 39: p. 269-281.

19. Gupta, S.B., et al., Ignition Characteristics of Methane-Air Mixtures at Elevated Temperatures

and Pressures. SAE Transactions Journal of Fuels and Lubricants, 2005. SAE 2005-01-2189.

20. Kitsopanidis, I. and W.K. Cheng, Soot Formation Study in a Rapid Compression Machine.

Journal of Engineering for Gas Turbines and Power, 2006. 128(4): p. 942.

21. Lee, D. and S. Hochgreb, Rapid Compression Machines: Heat Transfer and Suppression of

Corner Vortex. Combustion and Flame, 1998. 114(3): p. 531-545.

22. Donovan, M.T., et al., Demonstration of a free-piston rapid compression facility for the study of

high temperature combustion phenomena. Combustion and Flame, 2004. 137(3): p. 351-365.

23. Ribaucour, M., et al., A Rapid Compression Machine for Autoignition of Hydrocarbons. Journal

de Chimie Physique, 1992. 89: p. 2121-52.

Page 33: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

29

24. Mittal, G., A rapid compression machine - design, characterization, and autoignition

investigations, in Department of Mechanical and Aerospace Engineering. 2006, Case Western

Reserve University: Cleveland, OH. p. 179.

25. Minetti, R., et al., A rapid compression machine investigation of oxidation and auto-ignition of n-

Heptane: Measurements and modeling. Combustion and Flame, 1995. 102(3): p. 298.

26. Mittal, G. and C.-J. Sung, Aerodynamics inside a rapid compression machine. Combustion and

Flame, 2006. 145(1-2): p. 160.

27. Griffiths, J.F., et al., Temperature fields during the development of autoignition in a rapid

compression machine. Faraday Discussions, 2002. 119.

28. Würmel, J., et al., The effect of diluent gases on ignition delay times in the shock tube and in the

rapid compression machine. Combustion and Flame, 2007. 151(1–2): p. 289-302.

29. Tanaka, S., F. Ayala, and J.C. Keck, A reduced chemical kinetic model for HCCI combustion of

primary reference fuels in a rapid compression machine. Combustion and Flame, 2003. 133(4): p.

467-481.

30. Allen, C., et al., Application of a Novel Charge Preparation Approach to Testing the Autoignition

Characteristics of JP-8 and Camelina Hydroprocessed Renewable Jet Fuel in a Rapid

Compression Machine. Combustion and Flame, 2012. 159(9): p. 2780-2788.

31. Rui Xu, H.W., Med Colket, Tim Edwards, Thermochemical properties of jet fuels. National Jet

Fuel Combustion Program, 2015.

32. Edwards, T., C. Moses, and F. Dryer. Evaluation of combustion performance of alternative

aviation fuels. in 46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit. 2010.

33. Kyungwook Min, D.V., Anna Oldani, Tim Edwards, Tonghun Lee, AN IGNITION DELAY

STUDY OF CATEGORY C ALTERNATIVE AVIATION FUEL. ASME 2015 International

Mechanical Engineering Congress & Exposition, submitted, 2015.

34. Valco, D., et al. Autoignition Behavior of Synthetic Alternative Jet Fuels and Blends in a Rapid

Compression Machine. in 52nd AIAA Aerospace Sciences Meeting - AIAA Science and

Technology Forum and Exposition, SciTech 2014. 2014.

35. Kumar, K. and C.-J. Sung, An experimental study of the autoignition characteristics of

conventional jet fuel/oxidizer mixtures: Jet-A and JP-8. Combustion and Flame, 2010. 157(4): p.

676-685.

36. Vasu, S.S., D.F. Davidson, and R.K. Hanson, Jet fuel ignition delay times: Shock tube

experiments over wide conditions and surrogate model predictions. Combustion and Flame, 2008.

152(1-2): p. 125-143.

37. Dean, A.J., et al., Autoignition of surrogate fuels at elevated temperatures and pressures.

Proceedings of the Combustion Institute, 2007. 31(2): p. 2481-2488.

38. Kahandawala, M.S.P., et al., Ignition and Emission Characteristics of Surrogate and Practical

Jet Fuels. Energy & Fuels, 2008. 22(6): p. 3673-3679.

39. Allen, C., et al., JP-5 and HRJ-5 autoignition characteristics and surrogate modeling. Energy

and Fuels, 2013. 27(12): p. 7790-7799.

40. Allen, C., et al., Ignition behavior and surrogate modeling of JP-8 and of camelina and tallow

hydrotreated renewable jet fuels at low temperatures. Combustion and Flame, 2013. 160: p. 232-

239.

41. Dooley, S., et al., The experimental evaluation of a methodology for surrogate fuel formulation to

emulate gas phase combustion kinetic phenomena. Combustion and Flame, 2012. 159(4): p.

1444-1466.

42. Dooley, S., et al., The combustion kinetics of a synthetic paraffinic jet aviation fuel and a

fundamentally formulated, experimentally validated surrogate fuel. Combustion and Flame, 2012.

159(10): p. 3014-3020.

43. Pratt&Whitney, Evaluation of Amyris Direct Sugar to Hydrocarbon (DSHC) Fuel. Continuous

Lower Energy, Emissions and Noise (CLEEN) Program, 2014.07.

Page 34: IGNITION DELAY STUDY OF NEXT GENERATION ALTERNATIVE JET

30

44. Zhu, Y., et al., Ignition delay times of conventional and alternative fuels behind reflected shock

waves. Proceedings of the Combustion Institute, 2014.

45. Johnston, G., Alcohol to Jet.

46. Kumar, K. and C.-J. Sung, A comparative experimental study of the autoignition characteristics

of alternative and conventional jet fuel/oxidizer mixtures. Fuel, 2010. 89(10): p. 2853-2863.

47. Hui, X., et al., Experimental studies on the combustion characteristics of alternative jet fuels.

Fuel, 2012. 98: p. 176-182.