You are on page 1of 165

UNIVERSITY OF CINCINNATI

Mar. 1st, 2007


Date:___________________

JUN YANG
I, _________________________________________________________,
hereby submit this work as part of the requirements for the degree of:
Ph.D
in:
Environmental Engineering
It is entitled:
THERMAL DECOMPOSITION AND GROWTH OF SHORT ALKYLATED NAPHTHALENES

This work and its defense approved by:

Dr. Mingming Lu
Chair: _______________________________
Dr. Tim Keener
_______________________________
Dr. Soon-Jai Khang
_______________________________
Dr. San-Mou Jeng
_______________________________
_______________________________
Thermal Decomposition and Growth of Short Alkylated Naphthalenes

A dissertation submitted to the

Division of Research and Advanced Studies


of University of Cincinnati

In partial fulfillment of the


requirements for the degree of

DOCTORATE OF PHILOSOPHY (Ph.D.)

in the Department of Environmental Engineering


of the College of Engineering

2006

by

Jun Yang

B.S. (Environ. Eng.), Tsinghua University, P.R. China, 1996


M.S. (Environ. Eng.), Tsinghua University, P.R. China, 1999

Committee Chair:
Dr. Mingming Lu
Committee members:
Dr. Tim Keener
Dr. Soon-Jai Khang
Dr. San-Mou Jeng
ABSTRACT

As the most abundant category of PAHs in diesel fuel, alkylated naphthalenes are closely related

to PAH and soot formation during combustion. To understand the procedure of PAH and soot

formation so that the combustion technologies could be promoted to decrease PAH and soot

emission, it is important to investigate the thermal decomposition and growth of shortly-

alkylated naphthalenes. This study aims at a better understanding of: the mechanisms of short-

alkylated naphthalenes decomposition in high temperature conditions and PAHs/soot formation

therein; the comparison of the mechanisms with those of homolog bearing a single aromatic ring,

such as toluene and ethylbenzene. Experiments were conducted in a semi-isothermal tubular

flow reactor at different temperatures ranging from 550 to 1000ºC. Products including gaseous

hydrocarbons, aromatic hydrocarbons, CO, and soot were collected by the product collection

system. Then the products were identified and quantified. The major products of those

compounds at high temperatures and their profiles versus reaction temperature were presented.

The routes of reactant decomposition and product formation were discussed based on

experimental results and reaction mechanisms were postulated. In addition, the research has also

discussed: the effects of intrinsic chemical structure, such as the 1 vs. 2 positions of the methyl

group, and the resultant product distributions; the difference of 1-methylnaphthalene

decomposition in pyrolysis and oxidation conditions. The global kinetic analysis method was

employed to simulate reactant decomposition, and formation of some products and kinetic

parameters were obtained. The simulation results were in good agreement with experimental data.
ACKNOWLEDGEMENTS

I would like to acknowledge many people for helping me during my doctoral work. I would

especially like to thank my advisor, Dr. Mingming Lu, for her generous time and commitment.

Throughout my doctoral work she encouraged me to develop independent thinking and research

skills. She continually stimulated my analytical thinking and greatly assisted me with scientific

writing.

I am also very grateful for having an exceptional doctoral committee and wish to thank Dr. Tim

Keener, Dr. Soon-Jai Khang, and Dr. San-Mou Jeng, for their continual support and

encouragement.

I wish to thank Ming Chai for help me with my lab work. Also, I extend many thanks to my

colleagues and friends in the air quality group. Especially, I would like to thank Dr. Feng Shang

for his critical suggestion and powerful support with his knowledgeable background.

Finally, I'd like to thank my wife and my lovely daughter. Without their support and

encouragement, I could not finish my dissertation fluently.


1 TABLE OF CONTENTS

TABLE OF CONTENTS ..............................................................................................................I

LIST OF TABLES ....................................................................................................................... V

LIST OF FIGURES .................................................................................................................. VII

CHAPTER I INTRODUCTION ................................................................................................ 1

1.1 BACKGROUND AND MOTIVATION .................................................................................... 1

1.2 LITERATURE REVIEW ....................................................................................................... 6

1.2.1 Overall process of PAH and soot formation ........................................................... 6

1.2.2 Thermal decomposition of benzene and naphthalene............................................. 8

1.2.3 Thermal decomposition of alkylated benzene and PAHs ....................................... 9

1.3 RESEARCH OBJECTIVES ................................................................................................. 13

REFERENCE ................................................................................................................................ 15

CHAPTER II EXPERIMENTAL DESIGN............................................................................. 20

2.1 EXPERIMENT SETUP ....................................................................................................... 21

2.2 PRODUCT COLLECTION .................................................................................................. 25

2.3 FLOW CONTROL ............................................................................................................. 26

2.4 ANALYTICAL METHOD .................................................................................................. 27

2.4.1 Solution--GC/MS.................................................................................................. 27

2.4.2 Emission gas ......................................................................................................... 29

2.4.3 Soot ....................................................................................................................... 30

2.5 QUALITY ASSURANCE AND QUALITY CONTROL ............................................................ 31

i
REFERENCE ................................................................................................................................ 33

CHAPTER III PYROLYSIS OF METHYLNAPHTHALENES.......................................... 34

3.1 INTRODUCTION .............................................................................................................. 34

3.2 EXPERIMENTAL METHODS .............................................................................................. 35

3.3 RESULTS AND DISCUSSIONS ........................................................................................... 35

3.3.1 The pyrolysis of the 1- and 2-methylnaphthalenes ............................................... 35

3.3.2 The reactivity of 1- and 2-methylnaphthalene ...................................................... 40

3.3.3 The hypothesized reaction mechanism of 1-methylnaphthalene pyrolysis .......... 41

3.3.4 The formation of carbon doubling products ......................................................... 43

3.3.5 Soot particulate size distribution........................................................................... 44

3.3.6 Detection of gaseous products .............................................................................. 45

3.4 CONCLUSIONS ................................................................................................................ 45

REFERENCES .............................................................................................................................. 46

CHAPTER IV 2-METHYLNAPHTHALENE OXIDATION ................................................ 48

4.1 INTRODUCTION .............................................................................................................. 48

4.2 RESULTS AND DISCUSSION............................................................................................. 49

4.2.1 Products and their yields as functions of temperature .......................................... 49

4.2.2 The hypothesized reaction mechanisms................................................................ 52

REFERENCE ................................................................................................................................ 53

CHAPTER V GLOBAL KINETIC ANALYSIS OF METHYLNAPHTHALENE

PYROLYSIS................................................................................................................................ 55

ii
5.1 INTRODUCTION .............................................................................................................. 55

5.2 VALIDATION OF THE PLUG FLOW ASSUMPTION ............................................................. 58

5.3 BASIC MODELS OF REACTANT CONSUMPTION AND PRODUCT FORMATION ................... 63

5.4 RESULTS AND DISCUSSION............................................................................................. 65

5.4.1 Reactant Decay in Pyrolysis of Methylnaphthalenes............................................ 65

5.4.2 The influence of axial diffusion on the kinetic analysis ....................................... 68

5.4.3 Hydrogen Concentration Estimate in 1-Methylnaphthalene Pyrolysis................. 70

5.4.4 The Yield of Intermediates in Pyrolysis ............................................................... 72

5.5 CONCLUSION.................................................................................................................. 76

REFERENCE ................................................................................................................................ 77

CHAPTER VI ETHYLNAPHTHALENE PYROLYSIS........................................................ 80

6.1 INTRODUCTION .............................................................................................................. 80

6.2 EXPERIMENTAL SETUP ................................................................................................... 81

6.3 RESULTS AND DISCUSSION............................................................................................. 82

6.3.1 Experiment results ................................................................................................ 82

6.3.2 Postulated reaction mechanisms ........................................................................... 85

6.3.3 Global kinetic study on 1-ethylnaphthalene consumption.................................... 91

6.4 CONCLUSION.................................................................................................................. 92

REFERENCES .............................................................................................................................. 93

CHAPTER VII 1-PROPYLNAPHTHALENE PYROLYSIS ................................................ 96

7.1 INTRODUCTION .............................................................................................................. 96

7.2 EXPERIMENTAL DESIGN ................................................................................................. 97

iii
7.3 RESULTS AND DISCUSSION............................................................................................. 98

7.3.1 Experimental results.............................................................................................. 98

7.3.2 Postulated reaction mechanism........................................................................... 105

7.3.3 Kinetic analysis................................................................................................... 111

7.4 CONCLUSION................................................................................................................ 117

REFERENCE .............................................................................................................................. 118

CHAPTER VIII CONCLUSIONS AND RECOMMENDATIONS .................................... 120

8.1 CONCLUSIONS .............................................................................................................. 120

8.2 RECOMMENDATION...................................................................................................... 123

8.2.1 Further modeling work using Chemkin .............................................................. 123

8.2.2 Gaseous product analysis.................................................................................... 124

REFERENCE .............................................................................................................................. 124

APPENDIX A CALIBRATION CURVES FOR THE SYRINGE PUMP AND THE FLOW

METER...................................................................................................................................... 133

APPENDIX B EXPERIMENTAL DATA OF AVERAGE PRODUCT YIELDS .............. 136

APPDENDIX C EXPERIMENTAL DATA OF YIELDS OF RESIDUAL REACTANTS

AND SOME PRODUCTS USED IN GLOBAL KINETIC ANALYSIS ............................. 143

APPENDIX D MATLAB CODE TO ESTIMATE KINETIC PARAMETERS OF 2-

METHYLNAPHTHALENE FORMATION IN 1-METHYLNAPHTHALENE

PYROLYSIS.............................................................................................................................. 147

2 iv
LIST OF TABLES

TABLE 2-1 THE MAIN EXPERIMENTAL PARAMETERS FOR COMBUSTION STUDY .............................. 20

TABLE 2-2 PHYSICAL PROPERTIES OF ALKYLATED NAPHTHALENES 61 ........................................... 22

TABLE 2-3 TEMPERATURE PROGRAM BASED ON EPA METHOD 525.2............................................ 27

TABLE 2-4 WEIGHT OF THE SUBSTRATE WITH GREASE CHANGING WITH TIME (MG) ....................... 31

TABLE 3-1 DISTRIBUTION OF SOOT SIZE FORMED IN 1-METHYLNAPHTHALENE PYROLYSIS AT 950

AND 1000°C .......................................................................................................................... 44

TABLE 5-1. KINETIC STUDIES ON PYROLYSIS OF ORGANIC COMPOUNDS BASED ON PLUG FLOW

ASSUMPTION. ......................................................................................................................... 56

TABLE 5-2. CRITERIA FOR THE VALIDITY OF THE PLUG FLOW IDEALIZATION SUMMARIZED BY

CUTLER ET AL. ....................................................................................................................... 59

TABLE 5-3. CHARACTERISTIC TIMES AND NON-DIMENSIONAL NUMBERS OF THE EXPERIMENTAL

SETUP FROM 600 TO 1000°C. ................................................................................................. 62

TABLE 5-4 KINETIC PARAMETERS FOR REACTANT DECAY IN PYROLYSIS OF METHYLNAPHTHALENES

............................................................................................................................................... 66

TABLE 5-5 POSSIBLE REACTIONS IN 1-METHYLNAPHTHALENE DECOMPOSITION AND THEIR KINETIC

PARAMETERS. N DENOTES THE 1-NAPHTHYL RADICAL........................................................... 70

TABLE 5-6 KINETIC PARAMETERS OBTAINED FROM THE MODEL FOR ISOMERIZATION IN PYROLYSIS

OF METHYLNAPHTHALENES ................................................................................................... 75

TABLE 6-1 INITIAL EXPERIMENTAL CONDITIONS ............................................................................ 82

TABLE 6-2 HEATS OF REACTION (KJ MOL-1) FOR THE REACTANT CONSUMPTION PATHWAYS IN 1-EN

PYROLYSIS ............................................................................................................................. 87

v
TABLE 7-1 INITIAL EXPERIMENTAL CONDITIONS FOR 1 SECOND..................................................... 98

TABLE 7-2 INITIATION REACTIONS AND HYDROGEN ABSTRACTION REACTIONS AND THEIR

CALCULATED HEAT OF REACTION IN 1-PROPYLNAPHTHALENE PYROLYSIS........................... 106

TABLE 7-3 KINETIC PARAMETERS FOR REACTANT DECOMPOSITION AND PRODUCT FORMATION IN

PYROLYSIS OF 1-PROPYLNAPHTHALENE ............................................................................... 113

TABLE B-1 PRODUCT YIELDS IN 1-METHYLNAPTHALENE PYROLYSIS IN % CARBON INPUT .......... 136

TABLE B-2 PRODUCT YIELDS IN 2-METHYLNAPTHALENE PYROLYSIS IN % CARBON INPUT .......... 137

TABLE B-3 PRODUCT YIELDS IN 2-METHYLNAPTHALENE OXIDATION % CARBON INPUT .............. 138

TABLE B-4 PRODUCT YIELDS IN 1-EETHYLNAPTHALENE PYROLYSIS IN % CARBON INPUT ........... 140

TABLE B-5 PRODUCT YIELDS IN 1-PROPYLNAPTHALENE PYROLYSIS IN % CARBON INPUT FOR 1

SECOND ................................................................................................................................ 141

TABLE B-6 PRODUCT YIELDS IN 1-PROPYLNAPTHALENE PYROLYSIS IN % CARBON INPUT FOR 1.9

SECOND ................................................................................................................................ 142

TABLE C-1 1-METHYLNAPHTHALENE PYROLYSIS ........................................................................ 143

TABLE C-2 2-METHYLNAPHTHALENE PYROLYSIS ........................................................................ 144

TABLE C-3 1-ETHYLNAPHTHALENE PYROLYSIS ........................................................................... 145

TABLE C-4 1-PROPYLNAPHTHALENE PYROLYSIS FOR 1 SECOND .................................................. 146

TABLE C-5 1-PROPYLNAPHTHALENE PYROLYSIS FOR 1.9 SECOND ............................................... 146

vi
LIST OF FIGURES

FIGURE 1-1 CONCENTRATION AND PERCENTAGE OF EACH CATEGORY OF COMPOUNDS IN LOW

SULFUR DIESEL FUEL 3

FIGURE 1-2 A ROUGH PICTURE FOR SOOT FORMATION IN PREMIXED FLAMES 2 7

FIGURE 2-1 SCHEMATIC DIAGRAM OF THE REACTION SYSTEM 22

FIGURE 2-2 TEMPERATURE PROFILES IN THE QUART TUBULAR REACTOR 24

AT ATMOSPHERIC CONDITION WITH A HELIUM FLOW. 24

FIGURE 2-3 FLOW CHART OF EXPERIMENTAL SETUP FOR THE CALIBRATION OF THE FLOW METER 26

FIGURE 3-1 THE CHROMATOGRAM OF PRODUCTS IN 1-METHYLNAPHTHALENE PYROLYSIS 36

FIGURE 3-2 THE OVERALL TEMPERATURE SERIES OF PYROLYTIC REACTIONS OF 1-

METHYLNAPHTHALENE (TOP PANEL) AND 2-METHYLNAPHTHALENE (BOTTOM PANEL). “1MN”

REPRESENTS 1-METHYLNAPHTHALENE; “2MN” REPRESENTS 2-METHYLNAPHTHALENE;

“CARBON RECOVERY” REPRESENTS THE SUM OF RESIDUAL REACTANT, AROMATIC PRODUCTS

AND SOOT ON CARBON BASIS. 37

FIGURE 3-3 YIELDS OF AROMATIC PRODUCTS IN THE PYROLYSIS OF 1-METHYLNAPHTHALENE (TOP

PANEL) AND 2-METHYLNAPHTHALENE (BOTTOM PANEL). NUMBERS IN PARENTHESIS DENOTE

THE NUMBER OF ISOMERS. “1MN” REPRESENTS 1-METHYLNAPHTHALENE; “2-MN”

REPRESENTS 2-METHYLNAPHTHALENE. 39

FIGURE 3-4 THE SELECTIVE ION CHROMATOGRAM FOR C12H12 AND C12H10 ISOMERS IN 39

1-METHYLNAPHTHALENE PYROLYSIS. “EN” REPRESENTS ETHYLNAPHTHALENE AND “DMN”

REPRESENTS DIMETHYLNAPHTHALENE. 39

vii
FIGURE 3-5 THE SELECTIVE ION CHROMATOGRAM FOR C12H12 IN 2-METHYLNAPHTHALENE

PYROLYSIS. “EN” REPRESENTS ETHYLNAPHTHALENE AND “DMN” REPRESENTS

DIMETHYLNAPHTHALENE. 40

FIGURE 3-6 THE COMPARISON OF PYROLYTIC REACTIVITY OF THE TWO METHYLNAPHTHALENES.

THE SOLID LINES AND SOLID MARKERS REPRESENT 1-METHYLNAPHTHALENE DATA; THE

DASHED LINES AND HOLLOW MARKERS REPRESENT 2-METHYLNAPHTHALENE DATA. 41

FIGURE 3-7 HYPOTHESIZED PATHWAYS OF 1-METHYLNAPHTHALENE PYROLYSIS. 42

FIGURE 4-1 THE CHROMATOGRAM OF PRODUCTS IN 2-METHYLNAPHTHALENES OXIDATION AT

900°C. 1 BENZENE; 2 TOLUENE; 3 ETHYLBENZENE; 4 P-XYLENE; 5 PHENYLACETYLENE; 6

STYRENE; 7 BENZALDEHYDE; 8 BENZOFURAN; 9 INDENE; 10 BENZOFULVENE; 11

NAPHTHALENE; 12 2-METHYLNAPHTHALENES; 13 1-METHYLNAPHTHALENES; 14 C12H10S &

C12H12S; 15 ACENAPHTHALENE; 16 2-NAPHTHALDEHYDE; 17 FLUORENE; 18 C13H10S; 19

PHENANTHRENE; 20 ANTHRACENE 50

FIGURE 4-2 THE OVERALL TEMPERATURE SERIES OF 2-METHYLNAPHTHALENE OXIDATION. “2M”

REPRESENTS 2-METHYLNAPHTHALENE. 50

FIGURE 4-3 YIELDS OF AROMATIC PRODUCTS IN 2-METHYLNAPHTHALENE OXIDATION. NUMBERS IN

PARENTHESIS DENOTE THE NUMBER OF ISOMERS. “1MN” REPRESENTS 1-

METHYLNAPHTHALENE. 51

FIGURE 4-4 THE SELECTIVE ION CHROMATOGRAM FOR C12H12 IN 2-METHYLNAPHTHALENES

OXIDATION. “EN” REPRESENTS ETHYLNAPHTHALENE AND “DMN” REPRESENTS

DIMETHYLNAPHTHALENE. 52

FIGURE 4-5 HYPOTHESIZED PATHWAYS OF 2-METHYLNAPHTHALENE OXIDATION 53

viii
FIGURE 5-1 COMPARISON BETWEEN EXPERIMENTAL DATA (SOLID DIAMONDS) AND MODEL

PREDICTION FOR REACTANT CONSUMPTION IN 1-METHYLNAPHTHALENE PYROLYSIS (TOP

PANEL) AND 2-METHYLNAPHTHALENE (BOTTOM PANEL). THE THICK SOLID LINES ARE RESULT

OF THE NONLINEAR MODEL AND DASHED LINES GIVE THEIR 95% CONFIDENCE INTERVAL HALF-

WIDTHS. THE THIN SOLID LINES ARE RESULT OF THE LINEAR MODEL. 68

FIGURE 5-2 COMPARISON OF CALCULATED REACTION RATE CONSTANT OF REACTION 1 THROUGH 4

LISTED IN TABLE 5-5. “R1” REPRESENTS REACTION 1, “R2” DENOTES REACTION 2, AND SO ON.

72

FIGURE 5-3 THE REACTION MODEL FOR 2-METHYLNAPHTHALENE FORMATION IN 1-

METHYLNAPHTHALENE PYROLYSIS. X1M DENOTES THE CONCENTRATION OF 1-

METHYLNAPHTHALENE AND X2M DENOTES THE YIELD OF 2-METHYLNAPHTHALENE. 73

FIGURE 5-4 COMPARISON BETWEEN EXPERIMENTAL DATA AND MODELING RESULTS FOR THE

ISOMERIZATION PROCESS IN THE PYROLYSIS OF METHYLNAPHTHALENES. “Δ” DENOTES

EXPERIMENTAL DATA OF 2-METHYLNAPHTHALENE FORMATION FROM 1-

METHYLNAPHTHALENE PYROLYSIS, AND “×” DENOTES THOSE OF 1-METHYLNAPHTHALENE

FORMATION FROM 2-METHYLNAPHTHALENE PYROLYSIS. SOLID LINES DENOTE MODELING

RESULTS. DASH LINES MARK 95% CONFIDENCE INTERVAL HALF-WIDTHS OF PREDICTION. 75

FIGURE 6-1 TOTAL ION CHROMATOGRAM OF 1-ETHYLNAPHTHALENE PYROLYSIS AT 850 ºC 82

FIGURE 6-2 OVERALL TEMPERATURE SERIES OF 1-ETHYLNAPHTHALENE PYROLYSIS. “1-EN”: THE

QUANTITY OF UNCONSUMED 1-ETHYLNAPHTHALENE; “PAH”: THE TOTAL YIELDS OF ALL OF

AROMATIC PRODUCTS IDENTIFIED; “RECOVERY”: THE TOTAL RECOVERY, WHICH IS THE SUM

OF “1-EN” AND “PAH”. 83

ix
FIGURE 6-3 MAJOR PRODUCTS AND THEIR YIELDS IN 1-ETHYLNAPHTHALENE PYROLYSIS. “VN”

REPRESENTS VINYLNAPHTHALENE; “MN” REPRESENTS METHYLNAPHTHALENE; “ACEN”

REPRESENTS ACENAPHTHALENE. 84

FIGURE 6-4 PROPOSED REACTION MECHANISMS OF 1-ETHYLNAPHTHALENE PYROLYSIS. 88

FIGURE 6-5 THE ESTIMATED MAXIMUM YIELD OF 1-METHYLNAPHTHALENE. “RESIDUAL 1-EN”:

YIELDS OF UNCONSUMED 1-ETHYLNAPHTHALENE; “CONSUMED 1-EN”: QUANTITY OF

CONSUMED 1-ETHYLNAPHTHALENE, WHICH IS 1-“RESIDUAL 1-EN”; “1-VN + ACEN”: THE

SUM OF 1-VINYLNAPHTHALENE AND ACENAPHTHALENE; “MAXIMUM 1-MN”: THE ESTIMATED

MAXIMUM POSSIBLE YIELD OF 1-METHYLNAPHTHALENE IN 1-ETHYLNAPHTHALENE PYROLYSIS.

90

FIGURE 6-6 THE EXPERIMENTAL (HOLLOW DIAMOND POINTS) AND THEORETICAL RESULTS (CURVE)

OF 1-ETHYLNAPHTHALENE DECOMPOSITION. 92

FIGURE 7-1 A CHROMATOGRAM OF 1-PROPYLNAPHTHALENE PYROLYSIS AT 700°C. 1.TOLUENE, 2.

PHENYLETHYNE, 3. INDENE, 4. C10H10, 5. BENZOFULVENE, 6. NAPHTHALENE, 7. 2-

METHYLNAPHTHALENE, 8. 1-METHYLNAPHTHALENE, 9. 1-ETHYLNAPHTHALENE, 10. 1-

VINYLNAPHTHALENE, 11. 1,7-DIMETHYLNAPHTHALENE, 12. 2-VINYLNAPHTHALENE, 13. 14-

DIMETHYLNAPHTHALENE, 14. ACENAPHTHALENE, 15. 12-DIMETHYLNAPHTHALENE, 16. 1-

ALLYLNAPHTHALENE, 17. 1-PROPYLNAPHTHALENE, 18. ACENAPHTHENE, 19. 1-

ISOALLYLNAPHTHALENE, 20. FLUORENE, 21. C13H12, 22. PHENANTHRENE, 23. 1-

PHENYLMETHYL-NAPHTHALENE, 24. 1,1-BINAPHTHYL, 25. 1, 2’-DINAPHTHYLMETHANE, 26. 1,

2-DI-1-NAPHTHYLETHANE 99

FIGURE 7-2 THE YIELD OF RESIDUAL 1-PROPYLNAPHTHALENE AND TOTAL AROMATIC PRODUCTS AS

A FUNCTION OF TIME AND TEMPERATURE. “1PN” REPRESENTS 1-PROPYLNAPHTHALENE.

x
“CARBON RECOVERY” REPRESENTS THE SUM OF RESIDUAL REACTANT, AROMATIC PRODUCTS

AND SOOT ON CARBON BASIS. 101

FIGURE 7-3 MAJOR PRODUCT YIELDS OF 1-PROPYLNAPHTHALENE PYROLYSIS FOR 1 SECOND.

“1MN” REPRESENTS 1-METHYLNAPHTHALENE; “1EN” REPRESENTS 1-ETHYLNAPHTHALENE;

“1VN” REPRESENTS 1-VINYLNAPHTHALENE; “12DNE” REPRESENTS 1, 2-DI-1-

NAPHTHYLETHANE; “ACEN” REPRESENTS ACENAPHTHALENE; “12DNM” REPRESENTS 1, 2’-

DINAPHTHYLMETHANE. 102

FIGURE 7-4 THE MASS SPECTRA AND CHEMICAL STRUCTURES OF 1, 2-DI-1-NAPHTHYLETHANE AND

1, 2’-DINAPHTHYLMETHANE. 104

FIGURE 7-5 HYPOTHESIZED SCHEME OF REACTIONS HAPPENING IN 1-PROPYLNAPHTHALENE

PYROLYSIS. COMPOUNDS IN FRAMES ARE IDENTIFIED PRODUCTS AND OTHERS ARE RADICALS.

110

FIGURE 7-6 MODEL PREDICTIONS AND EXPERIMENTAL RESULTS FOR REACTANT DECAY IN 1-

PROPYLNAPHTHALENE PYROLYSIS. 113

FIGURE 7-7 MODEL PREDICTIONS AND EXPERIMENTAL RESULTS FOR PRODUCT FORMATION IN 1-

PROPYLNAPHTHALENE PYROLYSIS. 115

FIGURE 7-8 CALCULATED RETENTION TIME WHEN THE YIELDS OF 1- AND 2-METHYLNAPHTHALENE

AND 1-ETHYLNAPHTHALENE REACH THE HIGHEST VALUES AND THE CORRESPONDING YIELDS

FROM 550 TO 1000 ºC. 117

FIGURE A-1 THE CALIBRATION CURVE FOR THE SYRINGE PUMP 133

FIGURE A-2 CALIBRATION CURVES FOR THE FLOW METER IN HELIUM. “IN SYSTEM” REPRESENTS

THAT THE CALIBRATION CONDITION IS SAME AS IN PRACTICAL EXPERIMENTS EXCEPT HIGH

TEMPERATURES. “REAL DATA IN 1 ATM” REPRESENTS THAT THE FLOW METER IS CONNECTED

xi
TO THE BUBBLE METER DIRECTLY. “PROVIDED DATA” MEANS THE DATA PROVIDED BY THE

MANUFACTURER. “IN SYSTEM+NOZZLE” CONDITION IS SIMILAR TO THAT OF “IN SYSTEM”

EXCEPT THAT ONE NOOZLE IS CONNECTED TO THE BUBBLE METER. 134

FIGURE A-3 THE CALIBRATION CURVE FOR THE FLOW METER IN NITROGEN 135

xii
1 CHAPTER I INTRODUCTION

1.1 Background and Motivation

Growing attention has been given to polycyclic aromatic hydrocarbons (PAHs) in recent decades

due to the potential carcinogenicity and mutagenicity of some PAHs. For example,

benzo(a)pyrene has been proven to be related to lung cancer1. In coke manufacturing, there is a

high mortality rate due to skin and lung cancers in workers who are regularly exposed to high

atmospheric PAH concentrations which is indicative of the relationship between PAHs and

cancers.2 As a result, seven PAHs have been listed separately in the 1999 EPA National Toxics

Inventory. Those PAHs include benz(a)anthracene, benzo(a)pyrene, benzo(b)fluoranthene,

benzo(k)fluoranthene, chrysene, dibenz(a,h)anthracene, and indeno(1,2,3-cd)pyrene. Currently,

there are no legislations limiting the emission of PAHs due to their low emission concentration.

US EPA, however, suggested that when its state and local agencies report emissions of

polycyclic organic matter (POM), that they also provide detailed reports of 7-PAHs for future

consideration.3 Soot has been studied with PAHs because of two reasons:

(1) soot is formed from PAH precursors, and

(2) soot provides surface area for the adsorption and condensation of PAHs.

There are two types of PAH and soot sources: natural and anthropogenic. Naturally occurring

phenomena, such as forest fire and volcanic eruptions, produce both PAHs and soot.4 While

1
human activities related to the combustion of coal, oil, gas, wood and other fuels, such as mobile

combustion, coking processes, aluminum production, coal combustion, and domestic heating are

examples of the anthropogenic sources of PAHs.2 As human activity in manufacturing and

industry increases, more and more PAHs are emitted from these man-made sources. Mobile

sources are found to be the major source of ambient PAHs and account for 36% of the yearly

total emission in the US.5

It has been shown that the fuel type and combustion condition are two crucial elements in the

formation of PAHs and soot. For example, more PAHs are formed and released during the

combustion of diesel fuel than other common fuels. According to the study by Dunbar et al.6 on

the contributions of mobile sources of PAHs to urban air using a real-time PAH monitor in

downtown Boston, 39-65% of PAH emission from mobile sources was attributable to non-

passenger vehicles including buses and different types of trucks even though they only accounted

for six percent of total investigated vehicles. The measurement of PAH emissions in seven power

stations burning different fuels including coal, gasoline and diesel oil was conducted and the

results were compared to each other.7 It was pointed out that diesel-powered stations emitted

higher amounts of heavy PAHs than oil-fired stations. They concluded that combustion

conditions, the nature of the fuel and the type of furnace are the crucial factors which affect PAH

emission. It motivates us to know more details about the correlation between the chemical

compositions of diesel fuel and PAH and soot formation in combustion.

Short alklylated naphthalenes are important constituents of coal tars and petroleum products,

especially diesel fuel and jet fuels. Diesel fuel consists mainly of saturate and aromatic
2
hydrocarbons.8, 9 Paraffins, cycloparaffins, monoaromatics, diaromatics, polynuclear aromatics

and sulfur compounds are the major compound categories typically found in diesel fuel. In

Figure 1-1 the concentrations and percentages of each category of compounds in one sample of

low sulfur diesel fuel are shown.10 n-Alkanes, branched alkanes and cycloalkanes, which are

saturates, account for 27.90%, 53.87% and 7.72% of the total, respectively. The PAHs, most of

which are alkylated PAHs, include naphthalene, fluorene, phenanthrene, alkylated naphthalenes

and phenanthrenes, which constitute approximately 4% of the compounds identified in the diesel

fuel.,. Other studies have indicated that the volume fraction of PAHs in diesel fuels varies widely

from 5% to 30%.

100%
Alkylbenzenes, 6.6%
90%
Unidentified
80%
Alkylated PAHs, 3.7%
70% (31.6%)

60% PAHs, 0.26%


50%
Cycloalkanes, 7.7%
40% Identified
30% Branched alkanes,
(68.4%) 53.9%
20%

10%
n-Alkanes, 27.9%

0%

Figure 1-1 Concentration and percentage of each category of compounds in low sulfur diesel fuel 10

There are two sources of PAH emission in the combustion of fuels, unburned fuel11, 12 and the

combustion process in which both oxidation and pyrolysis may be involved. All high molecular

weight PAHs are formed during combustion, while light PAH constituents in fuel are burned

3
partially and the unburned part is expelled with emission gases.13 In the experimental study of

PAHs in engine soot by the method of isotopic tracing, it was observed that pyrene and

phenanthrene are formed from deuterated naphthalene added to fuel in the formed particulate

matter. In addition, residual deuterated naphthalene was identified.14

Usually, the combustion of aromatic hydrocarbons will produce more PAHs and soot than that of

aliphatic hydrocarbons and alcohols. This is the reason why diesel combustion produces more

PAHs and soot than the gas or oil combustion. Wang and Cadman observed that the order of the
15
sooting tendency in spray combustion is toluene > n-heptane > propanol. According to the

result of their study on soot formation in isothermal pyrolysis, the sooting tendencies of different

hydrocarbons at 1623 K relative to methane is: 1: 4: 7.6: 50: 7.4: 5.5: 4: 112: 91: 74 for methane:

ethylene: acetylene: diacetylene: benzene: toluene: p-xylene: naphthalene: anthracene: pyrene.16

It is understandable since the aromatic fragments formed in the combustion of aromatic

hydrocarbons provide a good source for heavy PAH and soot formation, while in aliphatic

hydrocarbon combustion those radical pieces have to be formed through different mechanisms,

such as dehydrogenation, cyclization and combination. In a study on the combustion process of

tetradecane in a direct injection diesel engine, Ciajolo et al. found that the addition of 10 vol.%

1-methylnaphthalene to diesel caused the increased formation of pyrolytic products including

light and heavy PAHs, acetylene and soot.17

To understand the procedure of PAH and soot formation from the combustion of fuels, it is

important to obtain insight into the chemical mechanisms of the combustion of different fuel

constitutents.18 Therefore, the thermal decomposition of short-alkylated naphthalenes, the most


4
important category of aromatic constituents in diesel, will be studied experimentally and

kinetically at different temperatures in this project. Since the formation of PAHs and soot is of

major interest and it is very closely related to the condition of pyrolysis or oxidative pyrolysis,

most of the experiments will be conducted in pyrolysis conditions. Products and residual

reactants will be collected and analyzed both qualitatively and quantitatively. The reaction

mechanism will be postulated and the method of global analysis will be employed to simulate

fuel decomposition and major PAH formation.19 Even though the traditional methods for

analyzing first- and second- order gaseous reactions are extensively used in many scientific

studies, it is difficult to use them to characterize reactions of practical interest which involve a

complex set of sequential and parallel unimolecular and bimoleculear reactions with the

consideration of the transport processes involved.19 Therefore, the traditional methods are not

used in this study. Alternatively, global kinetic data on reactions would be used as an initial step

to model practical combustion.20

Short alkylated naphthalenes are important aromatic constituents of coal tars and many heavy

oils such as diesel, aviation fuels and waste derived fuels. In combustion or at high temperatures,

they are relatively active since they can form resonance stabilized radicals and could facilitate

the formation of soot and heavy PAHs, even though the detailed mechanisms have not been

understood well. The study of thermal decomposition and further growth of short alkylated

naphthalenes will provide more information for the understanding of heavy PAH and soot

formation in the combustion practical fuels to diminish or eliminate pollutant emission. In

addition, the result of this study can be employed in chemical/petroleum industries to optimize

processes to minimize or maximize some specific products formed in high temperatures.


5
1.2 Literature Review

1.2.1 Overall process of PAH and soot formation

PAHs and soot are generally formed in incomplete combustion occurring in two conditions: (1)

In a flame where the ratio of air to fuel is low or the efficiencies of the mixing of air and fuel is

poor despite the high oxygen availability, and (2) post-flame where the temperature is lower than

that in flame.21 The process is largely determined by the thermo-chemistry of the reactivity of

precursors, either molecules or radicals, and is also related to fuel types and combustion

conditions.22, 23 PAH and soot formation usually involves four main steps roughly demonstrated

in Figure 12, 24 (1) formation of heavy PAHs of molecular weight 500 – 1000 amu, molecular

precursors of soot through the addition of C2, C3 or other small units to small molecules such as

benzene and PAH-PAH radical recombination and addition reactions; (2) nucleation or inception

of particles from heavy PAH molecules; (3) mass growth of particles by addition of gas phase

molecules and coagulation via reactive particle-particle collisions; (4) carbonization of particle

material. 2, 3

6
Figure 1-2 A rough picture for soot formation in premixed flames 2

Even though the general features of PAH and soot formation have already been determined, their

detailed formation procedures are different, and result from differences in fuels, combustor types,

combustion temperatures and fuel-air ratios. For example, it was found that the mass spectra of

soot produced from different normal alkane fuels displayed unique features that are different for

each parent fuel when the technology of laser-desorption time-of-flight mass-spectrometry was

employed to determine the chemical composition of soot.25 Therefore, it is necessary to study the

detailed mechanisms of PAH and soot formation from different hydrocarbons, especially

aromatics to decrease pollutant emission. In the following subsections some important studies on

PAH formation from the combustion of aromatic hydrocarbons will be briefly introduced.

7
1.2.2 Thermal decomposition of benzene and naphthalene

Aromatic hydrocarbons are getting more attention due to their higher tendency to form PAHs

and soot than common aliphatic fuels and their extensive existence in practical fuels. Knowledge

of benzene pyrolysis, the simplest aromatic compound, is fundamental in studying combustion of

aromatic hydrocarbons. Badger and Novotny26 observed that biphenyl was the major product in

benzene pyrolysis at 700°C. The amounts of p-terphenyl, m-terphenyl and triphenylene were

small, but still significant. Methane and ethylene were both recognized in the emission gas. They

concluded that the major process was the carbon-hydrogen fission to give phenyl radicals, which

formed biphenyl and terphenyl, even though the fission of the ring occurred. Richter et al.27 used

a molecular beam sampling coupled to mass spectrometry (MBMS) to study formation of PAHs

in premixed benzene flames based on the data of species and radicals obtained by Bittner et al..28

The cyclopentadienyl radical was found to be a key species for naphthalene formation.

Formation of heavier PAHs was proposed based on H abstraction and acetylene addition and

sometimes with the contribution of small PAHs.

Naphthalene is important since it is usually the most abundant pure PAH constituent in many

heavy oils and presents in the combustion products of various fuels in a variety of studies. It was

pointed out that the sooting tendency of naphthalene is far larger than other aliphatic or aromatic

compounds, including pyrene and anthracene at high temperatures.16 It was even chosen to

correlate soot emission with different fuels as an important component of many fuels and a

successful result was obtained.29 Badger et al. studied the formation of aromatic hydrocarbons at

700ºC via pyrolysis of [1-14C] naphthalene.30 Dimerization, which produces 1,1', 1,2', 2,2'-

binaphthyls, was the dominant mechanism. Perylene, 10,11-benzofluoranthene and 11,12-


8
benzofluoranthene were formed in smaller amounts. Benzo(j)fluoranthene and

benzo(k)fluoranthene were formed from the binaphthyls by cyclodehydrogenation, and further

carbon-hydrogen fission. The presence of small amounts of benzo(a)pyrene proved that the

cleavage of the ring system does occur to a small extent. The products of naphthalene oxidation

were far more complex than those in pyrolysis according to Griesheimer and Homann’s study. 31

The intermediates, resulting from the cleavage of rings, played important roles in oxidation,

which is different from those in pyrolysis experiments. Many other researchers noticed

cyclodehydrogenation in their studies and explained the experimental identification of different

biaryl isomers and of condensed PAHs containing a five-membered ring30-33.

1.2.3 Thermal decomposition of alkylated benzene and PAHs

As well as pure aromatics, alkylated aromatic hydrocarbons are important components of diesel

fuel, jet fuels, coal tars, and other heavy oils. It has been reported that methylated and ethylated

groups are the most abundant.34 Their chemical and biological properties may substantially differ

from those of their parent aromatics, the fact of which motivates more studies on their behaviors

in combustion. Due to the diversity of the radicals formed, the combustion of alkylated aromatic

hydrocarbons is much more complex than that of pure aromatic hydrocarbons. The pyrolysis

studies of alkylated benzenes with different chain lengths have provided insight into thermal

reactions during combustion of heavy oils.

There are two possible routes for the decomposition of toluene, the simplest alkylated aromatic

hydrocarbon: dehydrogenation on the methyl side chain giving rise to the benzyl radical and

cleavage of the aryl-alkyl carbon bond forming phenyl and methyl radicals. Using Knudsen cell
9
mass spectrometry, Smith35, 36 found that the first route is the major mechanism of initiation

reaction in toluene pyrolysis while the second one is unimportant at temperatures below 1400ºC.

Further reactions of the benzyl radical and other products occurred. Abundant stable products up

to C20H12 observed by other investigators37, 38 also presented in his products. Based on the results

of their model of the combustion of toluene, Lindstedt and Maurice39 also deduced that the major

path of toluene consumption is the one leading to the benzyl radical in plug flow reactors at

1200K and in diffusion flames. In the study of the oxidation of toluene, Emdee et al.40 pointed

out that the benzyl radical is still the major intermediate in the consumption of toluene through

the hydrogen or hydroxide radical abstraction of a methyl H. After comparing the thermal

reaction of PAHs from pyrolysis of different fuels in a tubular flow reactor at temperatures

between 700-1400ºC and 160KPa, Andreas Jess concluded that toluene can decompose at a

lower temperature and is more active than naphthalene and benzene 41.

The length of the alkyl chain has an important effect on the combustion of alkylbenzene. For

example, ethylbenzene pyrolysis is more complex than toluene’s. Many studies have shown that

the initial step in the pyrolysis of ethylbenzene is the C-C bond split in the side chain forming

benzyl and methyl radicals shown in route (1)42, 43


Even though Troe and his colleagues44

suggested that the route (2) dominates at higher temperatures (1250-1600K), others disagreed

with their suggestion and proposed route (1) as the dominant mechanism even at temperatures up

to 2080 K.45, 46

10
CH2
CH3
+ CH3
(1) (2)
CH
CH3
CH3
+ H

In the combustion of alkylated PAHs, the number and the configuration of aromatic rings, and

the position of the alkyl chain play important roles. Pyrolysis of methylanthracenes at

temperatures between 350 and 450 ºC for batch holding times up to 300 minutes was studied by

Smith and Savage.47 Three primary reaction pathways were identified: demethylation forming

anthracene, methyl addition forming dimethylanthracenes and hydrogenation resulting in methyl-

9,10-dihydroantracenes. Anthracene was the most abundant product in most reaction conditions,

which suggests that the first reaction pathway is the major route. It was proposed that for

pyrolysis of an alkylated PAH, logarithm of demethylation rate at 400ºC has a liner relationship

with the Dewar reactivity number Nts of the peripheral aromatic carbon atom bearing the alkyl

substituent: ln(rate) =3.7-7.1Nts. The Dewar reactivity number is a number proportional to the

difference between π electron energy of the aromatic reactant and that of the intermediate for a

family of aromatic substitution reactions. A similar conclusion was declared in Smith and

Savages’ earlier study.47

Tancell et al.48 found that vinylphenanthrenes were the major products in the combustion of 2-

and 3-ethylphenanthrene spiked into low-aromatic diesel fuel. Due to the high Dewar reactivity

numbers of the aromatic carbon connecting alkyl branch of those compounds, dealkylation is

insignificant or does not exist. Their results are consistent with Smith and Savage postulation47.

Similarly, in another combustion experiment of diesel spiked with 2-ethylnaphthalene and 2-


11
butylnaphthalene, no significant increase in the emission of naphthalene was observed. In fact,

with a high Dewar activity number, naphthalene is not supposed to be one of main products.

Instead, 2-vinylnaphthalene was found in products.

The comparison of pyrolysis of 1-methylpyrene and 1-ethylpyrene was done by Smith and

Savage49 and experiment conditions were similar to those in their other studies.34, 47
In 1-

methylpyrene experiments, the major pyrolysis products included pyrene and dimethylpyrene,

and hydrogenolysis was the dominant mechanism. The pyrolysis of 1-ethylpyrene led to pyrene

and methylpyrene as the major products, and pyrene was much more abundant. The rate of

hydrogenolysis for ethylpyrene was greater than that for methylpyrene because more radicals

were available for this compound.

The structure of the alkyl chain is also a critical parameter. For example, different from 2-

ethylphenanthrene, the flash vacuum thermolysis (FVT) of 4-ethynylpyrene resulted in only

pyrene.50 When the temperature reached 800 °C, the reactant was converted to pyrene completely.

Wornat et al.51-54 investigated the formation of ethynyl-PAHs and cyclopenta-fused PAHs (CP-

PAHs) and found that ethynyl-PAHs are formed by C2H2 additions to locations on the aryl

radical where cyclization is not possible. In other words, cyclization into a ring is preferred over

an ethynyl-PAH if it is possible. The point can be explained by the significantly lower energy

barrier for the cyclization reactions. Furthermore, cyclization to a ring from an ethynyl-PAH is

not a significant route in combustion due to its requiring significantly higher energy than the

aryl-vinyl cyclization reaction.51

12
In addition, pressure can sometimes affect the mechanism of pyrolysis, too. For example, when

reaction pressure is increased to a significantly higher degree, e.g. 150 bar, methane becomes a

major product in pyrolysis of 9-methylphenanthrene.55, 56

Comparatively, there is not so much work conducted on the combustion of alkylated

naphthalenes which are ubiquitous in various petroleum fuels and coal tars. Shaddix et al.

studied oxidation of 1-methylnaphthalene with the utilities of experimental and computational


57, 58
methods. The researchers observed that 1-naphthaldehyde, naphthalene, indene,

phenylacetylene, and benzene were, chronologically, the major aromatic intermediates at

atmospheric pressure and 1170 K. A mechanism of 1-methylnaphthalene oxidation was proposed

based on the intermediate species profiles. According to the mechanism, the resonantly stabilized

1-naphthylmethyl radical underwent radical-radical reaction to form 1-naphthaldehyde, which

decomposed to the 1-naphthyl radical. Naphthalene and the 1-naphthyl radical were oxidized to

the 1-naphthoxy radical, which further decomposed to the indenyl radical. Vinylbenzene, styrene,

phenyl radical, and acetylene were formed from further reaction and decomposition.

1.3 Research Objectives

As mentioned in the preceding section, aromatic hydrocarbons usually have more tendencies to

form PAHs and soot during combustion than aliphatic hydrocarbons. As important aromatic

compositions in practical fuels, short alklylated naphthalenes could promote PAH and soot

formation. However, there are not many investigations on the combustion of those particular

compounds as there are on benzene and alkylbenzene. The aim of the present work is to

13
investigate the characteristics of the thermal decomposition of short alkylated naphthalene both

experimentally and kinetically.

In pyrolysis experiments, information about fuel decomposition and the recombination of

fragments and radicals can be obtained without the involvement of active oxidizing agents

including OH and oxygen radicals in an oxidation condition. In other words, pyrolysis

experiments can provide some important information that cannot be provided or are not easily

observable in oxidation experiments. The results from pyrolysis combined with oxidation data

will provide clearer insight into the mechanisms of PAH and soot formation from the combustion

of fuels. Comparatively, oxidation of fuels is closer to the real combustion and the presence of

oxygen has three possible effects: (1) promotion of PAH production due to the increase of the

free radical pool; (2) selective PAH formation involving oxidation; (3) decline in PAH yields due

to oxidation.59 The dominance of those effects may vary in different combustion conditions,

which will produce different intermediates and products. Since thermal decomposition and PAH

formation are the main topic in this study, the oxidation experiment will be only considered for

one compound.

Therefore we plan to study pyrolysis of 1-methylnaphthalene, 2-methylnaphthalene, 1-

ethylnaphthalene, 1-proylnaphthalene and oxidation of 2-methylnaphthalene at different

temperature ranges between 550 to 1000 °C. We will investigate:

(1) major and minor products and their yields at different temperatures;

(2) change of yields of products with different reaction temperatures;

14
(3) the postulation of decomposition of fuels and formation routes of major products, i.e.

mechanisms;

(4) global kinetic analysis of reactant decay and formation of some major products.

And the detailed subtasks include:

(1) Pyrolysis of methylnaphthalenes

(2) Oxidation of 2-methylnaphthalenes

(3) Global kinetic analysis of methylnaphthalene pyrolysis

(4) Pyrolysis of 1-ethylnaphthalene

(5) Pyrolysis of 1-propylnaphthalene

Reference

1. Denissenko, M.; Pao, A.; Tang, M.; Pfeifer, G., Preferential formation of benzo[a]pyrene
adducts at lung cancer mutational hotspots in P53. Science 1996, 274, 430-432.

2. Bockhorn, H., Soot formation in combustion. Springer-Verlag: Berlin, Heidelberg and


New York, 1994; p 1-8, 165-192 and 253-274.

3. Mastral, A.; Callen, M., A review on polycyclic aromatic hydrocarbon (PAH) emissions
from energy generation. Environ. Sci. Technol. 2000, 34, 3051-3057.

4. NIKOLAOU, K.; MASCLET, P.; MOUVIER, G., SOURCES AND CHEMICAL-


REACTIVITY OF POLYNUCLEAR AROMATIC-HYDROCARBONS IN THE
ATMOSPHERE - A CRITICAL-REVIEW. SCIENCE OF THE TOTAL ENVIRONMENT 1984,
32 103-132.

5. Bjorseth, A.; Ramdahl, T., Handbook of polycyclic aromatic hydrocarbons: Emission


sources and recent progress in analytical chemistry. M. Dekker: New York, 1983; Vol. 2.

6. Dunbar, J. C.; Lin, C.-I.; Vergucht, I.; Wong, J.; Durant, J. L., Estimating the
contributions of mobile sources of PAH to urban air using real-time PAH monitoring. The
Science of the Total Environment 2001, 279, (1-3), 1-19.

7. Masclet, P.; Bresson, M.; Mouvier, G., PAH emitted by power stations and influence
15
of combustion parameters. Fuel 1987, 66, (1987), 556-562.

8. Song, C.; Hsu, C. S.; Mochida, I., Chemistry of diesel fuels. Taylor & Francis: New York,
2000.

9. Diesel fuel and exhaust emissions; World Health Organization: Geneva, 1996.

10. Liang, F. Composition and Formation Mechanism of Diesel Particulate Matter


Associated with Various Factors from A Non-road Diesel Generator. University of Cincinnati,
2006.

11. Rhead, M. M.; Pemberton, R. D., Sources of Naphthalene in Diesel Exhaust Emissions.
energy & fuels 1996, 10, 837-843.

12. Rhead, M.; Hardy, S., The sources of polycyclic aromatic compounds in diesel engine
emissions. Fuel 2003, 82, 385-393.

13. Williams, P.; Abbass, M.; Andrews, G., Diesel partculate emissions: the role of unburned
fuel. Combust Flame 1989, 75, 1-24.

14. Lombaert, K.; Le Moyne, L.; De Maleissye, J.; Amouroux, J., Experimental study of
PAH in engine soot by isotopic tracing Combust Sci Technol 2006, 178, (4), 707-728.

15. Wang, R.; Cadman, P., Soot and PAH production from spray combustion of different
hydrocarbons behind reflected shock waves Combust. Flame 1998, 112, (3), 359-370.

16. Tesner, P. A.; Shurupov, S. V., Soot Formation during Pyrolysis of Naphthalene,
Anthracene and Pyrene. Combust. Sci. Technol 1997, 126, 139-151.

17. Ciajolo, A.; D'Anna, A.; Barbella, R.; Bertoli, C., Combustion of Tetradecane and
Tetradecane/α-Methylnaphthalene in a Diesel Engine with regard to Soot and PAH Formation.
Combust. Sci. and Tech. 1992, 87, 127-137.

18. Brezinsky, K.; Hura, H.; Glassman, I., Oxidation/Pyrolysis chemistry as related to fuel
sooting tendencies. Energy & fuels 1988, 2, (4), 487-493.

19. Burnham, A. K.; Braun, R. L., Global Kinetic Analysis of Complex materials. energy &
fuels 1999, 13, (1), 1-22.

20. Ledesma, E. B.; Marsh, N. D.; Sandrowitz, A. K.; Wornat, M. J., Global Kinetic Rate
Parameters for the Formation of Polycyclic Aromatic Hydrocarbons from the Pyrolyis of
Catechol, A Model Compound Representative of Solid Fuel Moieties. Energy & Fuels 2002, 16,
1331-1336.

21. Mansurov, Z., Soot formation in combustion processes. Combustion, Explosion, and
Shock Waves 2005, 41, (6), 727-744.

16
22. Kiefer, J.; Mizerka, L. J.; Patel, M.; Wei, H., A shock-tube investigation of major
pathways in the high-temperature pyrolysis of benzene. J. Phys. Chem 1985, 89, 2013-2019.

23. Boehm, H.; Jander, H.; Tanke, D. In PAH growth and soot formation in the pyrolysis of
acetylene and benzene at high temperatures and pressures: Modeling and experiment, 12th
Symposium (International) on Combustion, Pittsburgh, 1998; The Combustion Institute:
Pittsburgh, 1998; pp 1605-1612.

24. Richter, H.; Howard, J. B., Formation of polycyclic aromatic hydrocarbons and their
growth to soot -- a review of chemical reaction pathways. Progress in Energy and Combustion
Science 2000, 26, 565-608.

25. Majidi, V.; Saito, K.; Gordon, A.; Williams, F., Laser-desorption time-of-flight mass-
spectrometry analysis of soot from various hydrocarbon fuels. Combust Sci Technol 1999, 145,
37-56.

26. Badger, G. M.; Novotny, J., Formation of aromatic hydrocarbons at high temperatures.
xii Pyrolysis of bezene. J. Chem. Soc. 1961, 3400-3402.

27. Richter, H.; Grieco, W. J.; Howard, J. B., Formation mechanism of polycyclic aromatic
hydrocarbons and fullerenes in premixed benzene flames. Combust Flame 1999, 119, 1-22.

28. Bittner, J. D.; Howard, J. B. In Composition profiles and reaction mechanisms in a near-
sooting premixed benzene/oxygen/argon flame, 18th Symposium (international) on Combustion,
1981; the combustion institute: 1981; pp 1105-1116.

29. Chin, J. S.; Lefebvre, A. H., Influence of fuel chemical properties on soot emissions from
gas turbine combustors. Combust Sci Technol 1990, 73, 479-486.

30. Badger, G. M.; Jolad, S. D.; Spotswood, T. M., The Formation of Aromatic
Hydrocarbons at High Temperatures: xx. The Pyrolysis of [1-14C]Naphthalene. Aust. j. chem.
1964, 17, 771-777.

31. Griesheimer, J.; Homann, K. H. In Large molecules, radicals ions, and small soot
particles in fuel-rich hydrocarbon flames Part II. Aromatic radicals and intermediate PAHs in a
premixed low-pressure naphthalene/oxygen/argon flame, 27th Symposium (International) on
Combustion, Pittsburgh, 1998; The Combustion Institute: Pittsburgh, 1998; pp 1753-1759.

32. Sarofim, A.; Longwell, J.; Wornat, M.; Mukherjee, J., The role of biaryl reaction in PAH
and soot formation. In Soot formation in combustion: mechanisms and models, Bockhorn, H., Ed.
1994; pp 485-496.

33. Masonjones, M. C.; Lafleur, A. L.; Sarofim, F., Biarene formation during pyrolysis of a
mixture of anthracene and naphthalene. Combust Sci Technol 1995, 109, 273-285.

34. Smith, C. M.; Savage, P. E., Reactions Of Polycyclic Alkylaromatics .5. Pyrolysis Of
Methylanthracenes. Aiche Journal 1993, 39, (8), 1355-1362.
17
35. Smith, R. D., A direct mass spectrometric study on the mechanism of toluene pyrolysis at
high temperature. J. Phys. Chem. 1979, 83, 2553-1562.

36. Smith, R. D., Formation of radicals and complex organic compounds by high-
temperature pyrolysis: the pyrolysis of toluene. Combust Flame 1979, 35, 179-190.

37. Brooks, C. T.; Cummins, C. P. R.; Peacock, S. J., Pyrolysis of toluene using a static
system. Trans Faraday Soc. 1971, 67, 3265-3274.

38. Errede, L. A.; Cassidy, J. P., Formation of aromatic hydrocarbons at high temperatures.
IX. Pyrolysis of toluene, ethylbenzene, propylbenzene, and butylbenzene, . J. Am. Chem. Soc.
1960, 82, 3653-3658.

39. Lindstedt, R. P.; Maurice, L. Q., Detailed kinetic modeling of toluene combustion.
Combust Sci Tech 1996, 120, 119-167.

40. Emdee, J.; Brezinsky, K.; GLASSMAN, I., A Kinetic-model for the Oxidation of
Toluene near 1200 K. J. Chem. Phys. 1992, 96, 2151-2161.

41. Jess, A., Mechanisms and kinetics of thermal reactions of aromatic hydrocarbons from
pyrolysis of solid fuels. Fuel 1996, 75, 1441-1448.

42. Robaugh, D. A.; Stein, S. E., Very-low-pressure Pyrolysis of Ethylbenzene,


Isopropylbenzene, and Tert-butylbenzene. Int. J. Chem. Kinet. 1981, 13, 445 -462.

43. Davis, H. G., Rate of formation of toluene from ethylbenzene. Int. J. Chem. Kinet. 1983,
15, 469-474.

44. Brouwer, L.; Muller-Markgraf, W.; Troe, J., Identification of Primary Reaction Products
in the Thermal Decomposition of Aromatic Hydrocarbons. In 20th Symp. Int. Combust., The
Combustion Institute: 1984; p 799.

45. Mizerka, L. J.; Kiefer, J. H., The High-temperature Pyrolysis of Ethylbenzene Evidence
for Dissociation to Benzyl and Methyl Radicals. Int. J. Chem. Kinet. 1986, 18, 363-378.

46. Pamidimukkala, K. M.; Kern, R. D., The High Temperature Pyrolysis of Ethylbenzene.
Int. J. Chem. Kinet. 1986, 18, 1341-1353.

47. Smith, C.; Savage, P., Reactions of Polycyclic Alkylaromatics:3. Struture and Reactivity.
AIChE Journal 1991, 37, 1613-1624.

48. Tancell, P.; Rhead, M.; Pemberton, R.; Braven, J., Diesel combustion of an alkylated
polycyclic aromatic hydrocarbon. FUEL 1996 75, 717-723.

49. Smith, C. M.; Savage, P. E., Reactions Of Polycyclic Alkylaromatics .4. Hydrogenolysis
Mechanisms In 1-Alkylpyrene Pyrolysis. Energy & Fuels 1992, 6, (2), 195-202.

18
50. Freccero, M.; Golfi, R.; Sarzi, A. M.; Rastelli, A., Origin of Pyrene under High
Temperature Conditions in the Gas Phase. The Pivotal Role of Phenanthrene. J. Org. Chem 1999,
64, 3861-3866.

51. Marsh, N. D.; Wornat, M. J. In Formation pathways of ethhynyl-subsituted and


cycloentafused polycyclic aromatic hydrocarbons, Proceedings of Comubsiton Institute, 2000;
2000; pp 2585-2592.

52. Wornat, M. J.; Ledesma, E. B., C16H10 etynyl-substituted polycyclic aromatic


hydrocarbons from the pyrolysis of coal, coal volatiles, and anthracene. Polycylic Aromatic
Compounds 2000, 18, 129-147.

53. Wornat, M. J.; Vriesendorp, F. J. J.; Lafleur, A. L.; Plummer, E. F.; Necula, A.; Scott, L.
T., The identification of new ethynyl-substituted and cyclopenta-fused polycyclic aromatic
hydrocarbons in the products of anthracene pyrolysis. Polycylic aromatic compounds 1999, 13,
221-240.

54. Marsh, N. D.; Wornat, M. J.; Scott, L. T.; Necula, A.; Lafleur, A. L.; Plummer, E. F., The
identification of cyclopenta-fused and ethynyl-subsitituted polycyclic aromatic hydrocarbons in
benzene droplet combustion products. Polycylic aromatic compounds 2000, 13, 379-402.

55. Behar, F.; Budzinski, H.; Vandenbroucke, M.; Tang, Y., Methane Generation from Oil
Cracking: Kinetics of 9-Methylphenanthrene Cracking and Comparison with Other Pure
Compounds and Oil Fractions. Energy & fuels 1999, 13, 471-481.

56. Lorant, F.; Behar, F.; Vandenbroucke, M.; Mckinney, D. E.; Tang, Y., Methane
generation from methylated aromatics: kinetic study and carbon isotope modeling, Energy Fuels.
Energy & fuels 2000, 14, 1143-1155.

57. Shaddix, C. R.; Brezinsky, K.; Glassman, I., Oxidation of 1-Methylnaphthalene. In 24th
Symp. Int. Combust., The Combustion Institute: 1992; pp 683-690.

58. Shaddix, C. R.; Brezinsky, K.; Glassman, I., Analysis of Fuel Decay Routes in the High-
Temperature Oxidation of 1-Methylnaphthalene. Combust. Flame 1997, 108, 139-157.

59. Ledesma, E. B.; Kalish, M. A.; Nelson, P. F.; Wornat, J. J.; Mackie, J. C., Formation and
fate of PAH during the pyrolysis and fuel-rich combustion of coal primary tar. Fuel 2000, 79,
1801-1814.

19
2 CHAPTER II EXPERIMENTAL DESIGN

There are three typical categories of experimental conditions for the study on the formation of

aromatic compounds60, listed Table 2-1. The study of stabilized flames, including premixed

flame and diffusion flame, just examines the macroscopic, phenomenological parameters that

affect combustion such as flame type and temperature. Yet, not many details about microscopic,

chemical processes could be obtained from flame experiments.18 The experiments on shock tubes

usually are finished in a very short time and only emphasize the decomposition of fuel.

Contrastly, reactor experiments can provide insight into the decomposition of fuels and further

PAH and soot formation, which makes it a plausible choice for this study.

Table 2-1 The main experimental parameters for combustion study

Maximum Pressure(bars) Retention


Temperature(K) time(s)
Reactors 1300 1-40 0.1-100
Shock tubes 2500 1-40 10-4
Stabilized flames 2000 <= 1 10-3

All of the reactor experiments had been conducted in a bench tubular reactor at atmospheric

pressure. In every experiment, a certain quantity of reactant was put/injected and vaporized in a

heated vaporizer, and the vapor was taken away by carrier gas. The mixed gas then entered a

semi-isothermal reactor and the reactant was oxidized or pyrolyzed for one second. Products,

including soot and Polycyclic Aromatic Compounds (PACs), and residual reactant leaving the

reactor with the carrier gas were collected by a product collection system. Once the experiment

20
was finished, the reactor and the product collection system were rinsed with dichloromethane

(DCM). After being extracted from soot, PAHs and other products were identified and quantified

using a gas chromatography /mass spectrometry (GC/MS). Soot, if formed, was determined

gravitationally. In the succeeding sections, the experiment setup, product collection and analysis

are to be introduced in detail.

2.1 Experiment Setup

The schematic diagram of the experimental system is shown in Figure 2-1. The system consists

of three major parts: a vaporizer, a reactor, and a sample collection assembly. In the experiments,

a reactant was either injected by a syringe pump or put directly into the glass vaporizer which

was pre-heated to 200 ºC. It melted (if it was originally solid) and/or was vaporized gradually.

The syringe pump was calibrated before the experiments and the calibration curve is shown in

Table A-1. The vapor was delivered into the reactor by the carrier gas. The carrier gas was

prepurified helium (99.995%, Wright Brothers, Inc.) for the pyrolysis experiments, and 3470

ppm oxygen in nitrogen (Wright Brothers, Inc.) for the oxidation experiments. The inlet, the tube

reactor and the collection system were connected by stainless steel Swagelok fittings.

21
Reactant
input Furnace with
iso-thermo control
Quench
gas
Carrier
Heated
Gas GF filter
Vaporizer

Flow rate
control To hood

Dual trap
Quartz system
tubing
ice bath

Figure 2-1 Schematic diagram of the reaction system

1). The vaporizer is a three-way glass container connected to the reactor and a carrier gas

transfer line. The vaporizer and the transfer line were wrapped separately by two

Barnstead/Thermolyne flexible electric heating tapes. A double proportional voltage power

controller (Electrothermal MC240X1) was employed to control the temperature of the

heating tapes and the transfer line so that the temperature inside the vaporizer was kept at

approximately 200°C during the duration of the experiments. That temperature is higher

than the melting points of short alkylated naphthalenes, but lower than their boiling points.

Those physical properties are listed in Table 2-2. At that temperature, the reactant was

gradually vaporized without decomposition and the gasified reactant was delivered into the

reactor by carrier gas. It is important to heat the entire reactor and even the connector

between the vaporizer and the reactor to prevent reactant vapor from condensing, which

will decrease reactant input and fail the experiment.

Table 2-2 Physical properties of alkylated naphthalenes 61

22
Reactant Chemical Molecular Melting Point / Boiling Point /
Structure Weight K K
1-methylnaphthalene CH3 142 244±9 515±7

2-methylnaphthalene CH3 142 307±1 514±1

1-ethylnaphthalene 62 CH3 156 258.27 531.8

1-propylnaphthalene CH3 170 264.55 546±3

2). The reactor was a 0.965 m quartz tube with 1.7 cm inner diameter put in a tube furnace

(Thermolyne 79400). The furnace was equipped with a digital temperature control to heat

the reactor up electronically to desired temperatures. The ends of the tube were fire-

polished in case they crashed promptly while the system was assembled. The furnace

temperature profiles shown in Figure 2-2 were obtained using a long electric temperature

probe (Omega molded quick-connect probe KQSUP-14G-12) to present the axial

temperature distribution in the reactor from 700 to 900ºC. The temperatures in the middle

24 inches of the reactor, i.e. the reaction zone, are consistent and within ±10ºC of the set

point. It proved that the combination of the furnace and the middle section of quartz tube

can be used as a semi-isothermal reactor. The length of the reactor’s effective section is 24

inches.

23
1000

800

Temperature ( ˚C)
600

900
400 800
700

200
0 5 10 15 20 25 30 35
Reactor length (inch)

Figure 2-2 Temperature profiles in the quart tubular reactor


at atmospheric condition with a helium flow.

3). The product collection assembly, including one piece of glass fiber filter paper (Whatman

GF/F) and a dual-trap glass impinger, was designed to collect all of the soot and products

except some minor volatile or gaseous products, e.g. methane, ethylene, carbon monoxide,

and carbon dioxide. The filter paper was set in a fitting which connected the impinger and

the reactor and captured most of the soot formed as well as part of the PAH products. The

impinger was filled with solvent and placed in an ice bath to approach a higher collection

efficiency of PAH products. Dichloromethane (DCM, Acros HPLC grade, 99.9%), which

has a low boiling point of 40ºC, was selected as the solvent. Methylene chloride's volatility

and ability to dissolve a wide range of organic compounds makes it an ideal solvent.

Compressed room temperature air blew across the end of the quartz tubing reactor to cool

emission gas quickly and quench any further reactions immediately. The measured gas

temperature in the quenched fitting was less than 80°C.

24
2.2 Product Collection

After an experiment was finished, clear DCM was used to rinse the parts of the system that

touched the reactant and products, including the vaporizer, the quartz tube, the sample collection

assembly, and fittings. The used rinse liquid was then combined with the DCM solution in the

solvent trap for further extraction. The final solution was vacuum filtered by a PTFE filter paper

(Alltech PTFE Pre-Cut Membrane) put in a glass microanalysis vacuum filter holder

(Fisherbrand 09-753G) to remove soot. During filtration, all of the containers that had contact

with the final solution were rinsed using DCM three times, and the rinse liquid was filtered again

and combined with the clean solution to ensure that all of the solvable products were in the final

extracted solution. The volume of the extracted solution was determined by a graduated cylinder.

Soot was formed and collected at three locations: the glass fiber filter paper, the inner surface of

the quartz tubing and in the solution in the impinger. It was assumed that the collection

efficiency of the filter paper and the impinger was close or equal to 100% and no soot was in the

emission gas. Soot in the solution was separated by filtration, while that sticking to the quartz

tubing was scraped by a metal ring with sharp edge, then wiped with small pieces of Kimwipes

and rinsed with DCM. The mass of soot was determined by the difference of the filter and

Kimwipes pieces before and after being used.

In some experiments, emission gas was collected by SKC sampling bags connected to the end of

the impinger and was analyzed by different gas analyzers later. The gas sampling was usually

finished in four to six minutes depending on the specific experiment conditions. The volume of

25
the emission gas was larger than that of the pure carrier gas, considering the significant

vaporization of DCM, and increased approximately 20%.

2.3 Flow Control

The designed retention time, which is 1 second in most experiments, and the length of the

reaction zone, 24 inch or 0.61 m, decided the flow rate at a specific temperature. The flow rate

was controlled and monitored by a flow meter (Matheson, E11C071E501) located between the

carrier gas cylinder and the vaporizer. The pressure at the location was close to one atmosphere

and there was no vapor of reactant and DCM contaminates the flow meter. The meter was

calibrated using pure helium and nitrogen to approach accurate flow control.

The calibration of the flow meter was conducted with a bubble meter connected to the exit end of

the experimental system, which is shown in Figure 2-3. In different calibration conditions, the

volume of carrier gas was measured in a predetermined duration and the flow rate was calculated

by the gas volume divided by time. After a series of experiments, the relationships between the

flow meter reading and the corresponding flow rate with different carrier gases were decided and

shown in Table A-2 and Table A-3. It is found that the real flow rate is about 20% above the

value provided by the manufacturer for the same flow meter reading.

Gas Flow Different Calibration Bubble


Cylinder Meter Conditions Meter

Figure 2-3 Flow chart of experimental setup for the calibration of the flow meter

26
For helium, three conditions were tested: cylinder and flow meter only, which donates real

atmospheric pressure condition, experimental condition without heating, experimental condition

without heating and with a nozzle installed at the end of experimental system. The last one was

used to simulate the condition of experiments with emission gas collection using a sampling bag.

For nitrogen, only the experimental condition without nozzle was checked since emission gas

was not collected and analyzed.

2.4 Analytical Method

2.4.1 Solution--GC/MS

The extracted solution was analyzed using Varian Saturn 2200 GC/MS system, which was

equipped with a CP-Sil 8 CB Low Bleed/MS capillary column equivalent to a J&W DB5 column.

EPA method 525.2, which is the most commonly used method for semi-volatile compound

analysis including EPA 16 PAHs63, was employed with slight modification. The column oven

temperature program is set and shown in Table 2-3:

Table 2-3 Temperature program based on EPA method 525.2

Temp (°C ) Rate (°C /min) Hold (min) Total (min)


40 1.5 1.5
200 10.0 0.0 14.5
270 5.0 0.0 28.5
300 10.0 10.0 41.5

27
The Pressure Pulse of 40psi which was held for 0.80 minute was used for better sample

introduction. The column flow was kept constant at 1.2 ml/min after that point and hold for the

remaining duration.

In the practical analysis, the start temperature of the column oven was set to 40°C instead of the

original value, 70°C, to extend the analysis ability to detect toluene, and other alkylated benzenes.

Those compounds presented in products and cannot be measured using the original EPA 525.2

method.

The analytical method was sometimes modified to separate isomers, such as

dimethylnaphthalenes. The retention times of those isomers are usually very close. The

temperature ramp rate sometimes needed to be adjusted to enlarge the difference of the isomers’

retention times and avoid or decrease the overlap of their peaks. The molecular formulas and

possible structures of PAHs were identified by comparing their mass spectra with standard

spectra in the NIST library. For the compounds that have isomers with similar spectra, samples

of those isomers which are commercially available were bought and analyzed using the same

GC/MS method. The compounds were finally identified through the comparison of their

retention times with those of the standards. We referred to related research to determine those

compounds that were not commercially available. A series of standard mixes that contained most

of the products were prepared and used to quantify different products. The quantities of those

compounds that had no corresponding standards were determined by the response factors of their

isomers or other compounds with close structures. More details about the analysis of

hydrocarbon using the GC/MS could be obtained from Dr. Fuyan Liang’s dissertation.10

28
Due to the limitation that GC/MS can only separate and identify the aromatic compounds

containing 22 or fewer carbon atoms, HPLC was employed to identify those heavier PAHs

possibly present in products.

2.4.2 Emission gas

Three approaches to analyze emission gas including two GCs with different detectors and one

FTIR were conducted.

1. HP 5890 series II equipped with a dual column HP 10 ft BX-45/60 mesh and HP 6 ft

HAYESEP Q 80/100 mesh, and TCD detector

The equipment is located in the Engineering Research Center, University of Cincinnati, and is

designed to measure the concentration of CO2 and methane in gas samples. Methane was found

in the emission gas of pyrolysis of 1-methylnaphthalene at 900ºC and its portion of carbon input

is less than 0.5%.

2. Multiple gas analyzer based on Fourier transform infrared (FTIR) Spectroscopy

The instrument is located in the Center Hill Research Center, University of Cincinnati. FTIR

works based on the fact that most gas molecules except noble gases and homonuclear diatoms,

which are infrared inactive, absorb infrared radiation with a characteristic, molecule-specific

absorption spectrum. Through the location and magnitude of these absorptions gas species can be

identified and quantified.64

29
3. GC-FID

The instrument is located at the Ohio University and was equipped with a cryogenic pre-

concentration system. Since the concentration of DCM in the gas sample was too high (reached

20%) and it contaminated the pre-concentration system, only one sample was analyzed by the

instrument.

In order to simplify tasks, the products escaping with the final emission gas were neglected and

emission gas was not collected in most of the pyrolysis experiments. In the oxidation

experiments, further detection of emission gas may be needed on account of the complexity of

the oxidation products.

2.4.3 Soot

An analytical balance (Sartorius CP 225D) was employed to detect the soot mass gravitationally.

The mass of soot on the filter papers was determined by the mass difference of the dry papers

before and after experiments. Those wiper pieces were measured before and after experiments,

also. The difference of the total mass of the wipes was the mass of the soot sticking to the reactor.

The total amount of soot is assumed to be the sum of that on the glass filter, of that extracted

from solution and of that on the surface of the reactor.

The size distribution of soot in two pyrolysis experiments was measured by a personal eight-

stage cascade impactor SKC 225-50-001. The impactor was connected directly to the end of the

reactor with a designed flow rate of 2 liter per minute. Different from other experiments, the

experiment time usually lasted for less than one minute due to the overload of the cascade

30
impactor. The product collection system was not used in that condition. During the sampling,

silica grease was smeared on the surface of metal collection substrates. Since silica grease

contains a small portion of volatile constituents, substrates are put aside for a few days before

being used. In Table 2-4 the change of weight of a substrate smeared with grease in 65 hours is

shown.

Table 2-4 Weight of the substrate with grease changing with time (mg)

0 hr 43rd hr 65th hr

165.73 164.68 164.67

In addition, some volatile products were condensed or adsorbed on the surfaces of substrates

during sampling. Those compounds vaporized gradually and disturbed the determination of soot

mass. Thus substrates were weighed 24 hours after the experiment was finished.

2.5 Quality Assurance and Quality Control

Before the experiments, some preliminary tests were performed for quality assurance. A blank

run at 1000°C was conducted using the above system without introducing a reactant to detect the

existence of any possible contaminants in the system. The reliability of the experimental system

was proven by the fact that no contaminants were found in either the DCM solution or the

emission gas. The product recovery test at 200ºC, which is not high enough for the reactant to

decompose, was calculated to check the collection efficiency and whether there was any leakage

in the experimental system. 97.7% recovery was obtained. Repeated experiments were conducted

31
in every experimental condition to reduce possible deviation in some experiments. Other QA-QC

steps conducted in the analytical procedure included:

1. Sample duplicates were used to reduce differences that might be obtained during the

analytical processes;

2. Sample containers, sampling equipment, and laboratory glassware were cleaned and

stored to ensure that there was none or an insignificant existence of contamination from

external sources;

3. Analytical mix solutions sealed in vials were stored in refrigerators and were reprepared

after six months to one year or they become deteriorated. Simple fresh standard solution

is prepared to verify the quality of analytical solution.

Relying on the above experimental design we conducted a series of experiments of pyrolysis and

oxidation of alkylated naphthalenes over a selected temperature range from 550 to 1000°C with

an increment of 25/50°C, and products and their yields at every experiment conditions were

obtained. Each run lasted for four minutes, plus an additional two minutes to flush the residual

reactant and products in the reactor. Multiple experiments were conducted in every experimental

condition to ensure accuracy. As an important parameter of evaluating the quality of experiment,

carbon recovery was calculated by adding all of the carbon in the residual reactant, aromatic

products and compared to the sum of that in the input reactant. For the convenience of

comparison, all quantities were expressed in the percentage of carbon input. Even though soot

not only consists of carbon atoms but also contains up to 10% moles of hydrogen65, we treated

soot as pure carbon since the percentage and the molecular weight of hydrogen are much lower

than those of carbon.

32
Reference

1. Vovelle, C.; Delfau, J. L., Formation of aromatics in combustion systems. In Pollutants


from combustion : formation and impact on atmospheric chemistry, Vovelle, C., Ed. Kluwer
Academic Publishers: Boston, 2000.

2. Brezinsky, K.; Hura, H.; Glassman, I., Oxidation/Pyrolysis chemistry as related to fuel
sooting tendencies. Energy & fuels 1988, 2, (4), 487-493.

3. NIST Chemistry WebBook, NIST Standard Reference Database Number 69, June 2005
Release. In 2005.

4. Weast, R. C.; Grasselli, J. G., CRC Handbook of Data on Organic Compounds. second
ed.; CRC Press, Inc.: Boca Raton, FL, 1989; Vol. 1.

5. Eichelberger, J. W.; Munch, J. W.; Shoemaker, J. A., EPA 525.2 Determination of


organic compounds in drinking water by liquid-solid extraction and capillary column gas
chromatography/mass spectrometry. In Varian Environmental Application Manual.

6. Liang, F. Composition and Formation Mechanism of Diesel Particulate Matter


Associated with Various Factors from A Non-road Diesel Generator. University of Cincinnati,
2006.

7. El-kady, A. M.; Jeng, S.-M., Experimental Investigation of Temperature and Species


Concentration Characteristics of Swirling Spray Combustion. In 43rd Aerospace Sciences
Meeting & Exhibit, Reno, Nevada, 2005.

8. Seinfeld, J. H.; Pandis, S. N., Atmospheric Chemistry and Physics: From Air Pollution to
Climate Change. John Wiley & Sons: New York, 1997.

33
3 CHAPTER III PYROLYSIS OF

METHYLNAPHTHALENES

3.1 Introduction

Studies on the oxidation of 1-methylnaphthalene by Shaddix et al. indicated that the abstraction

of benzylic hydrogen is the major reaction mechanism at approximately 1200K1, 2. Pitsch3 has

provided a detailed kinetic reaction mechanism describing ignition and oxidation of 1-

methylnaphthalene and verified it using the above mentioned experimental data. Goos et al. has

also studied the early soot formation of 1-methylnaphthalene at high pressures based on shock

tube experiments4, which used Pitsch’s gas phase models. The results demonstrated the

importance of the resonance stabilized naphthylmethyl radicals. However, the products of carbon

doubling were not reported, whereas these types of compounds have been identified as major

products in the pyrolysis of toluene, indene and cyclopentadiene. A study of 2-

methylnaphthalene pyrolysis at 950°C has identified several carbon doubling products of

approximately 1.53% in total without reaction mechanisms and has reported on major products,

but the reaction mechanisms and major products were not provided.5 In addition, there has not

been a temperature series on the pyrolysis of methylnaphthalene, which are generally more

effective in investigating the PAH and soot formation mechanisms of targeted materials than are

oxidative studies, as the latter can result in more product fragmentation rather than product

growth.

34
This chapter aims at better understanding the PAH and soot formation mechanisms of 1- and 2-

methylnaphthalenes, the interpretation of the lack of major carbon doubling products, and the

differences of methylnaphthalenes with other compounds bearing benzylic structures, such as

toluene and indene. In addition, the chapter also discusses the effects of intrinsic chemical

structure, such as the 1 vs. 2 positions of the methyl group, and the resultant product distributions.

Experimental results of PAH and soot formation from the pyrolysis of 1 and 2-

methylnaphthalenes are presented, and the product formation pathways are hypothesized.

3.2 Experimental methods

The pyrolytic experiments were performed for both 1-methylnaphthalene (Acros, 97%, a liquid)

and 2-methylnaphthalene (Acros, 99%, a solid) over the temperature range of 800°C to 1000°C

with an increment of 25 or 50°C. The reactant concentration is approximately 500 ppm for

pyrolysis. The retention time is set to 1 second for further comparison.

3.3 Results and discussions

3.3.1 The pyrolysis of the 1- and 2-methylnaphthalenes

One Chromatogram of products from 1-methylnaphthalene pyrolysis at 950ºC is shown in Figure

3-1 and parts of products are labeled. Those unlabelled peaks are signals of septum bleed where

species containing silicone are involved.

35
2-Methylnaphthalene

1-Methylnaphthalene
2.0 RIC all 1m950_5_3.SMS

Naphthalene
MCounts

Acenaphthalene
1.5

C12H10(4) and C12H12(9)

Phenanthrene
1.0

Fluorene
Indene

Benzofulvene

Anthracene
0.5
Toluene

Styrene

0.0

5 10 15 20 25 minutes 30

Figure 3-1 The chromatogram of products in 1-methylnaphthalene pyrolysis

The temperature series of 1- and 2-methylnaphthalene pyrolysis is shown in Figure 3-2. Average

values are used for each temperature, and the error bar represents the actual maximum and

minimum values. At 800ºC, less than 20% of 1-methylnapththalene is consumed (as represented

by residual 1-methylnapththalene), while at 1000ºC, almost all has been decomposed with soot

accounting for 50% of the reactant input. This indicates that the selected temperature range is

suitable for studying the reactivity of 1-methylnaphthalene. The term “aromatic products” refers

to all the solvent extractable products, which increases from 5% at 800ºC to approximately 56%

at 950ºC and starts to decrease with temperature increase as they are increasingly converted to

soot. The total recovery of products and unconsumed reactant remains high (more than 80%)

until the temperature reaches 900ºC. At this point, between 900-975ºC, the recovery of products

and reactant decreases, but resumes with further temperature increase. This trend, we believe, is

related to the dominant pyrolytic mechanisms in each temperature range, which will be further

36
discussed subsequently. The temperature series of 2-methylnaphthalene pyrolysis in the bottom

panel of Figure 3-2 are analogous to those of 1-methylnaphthalene.

100

80
Yield ( % carbon input)

60

residual 1MN
40 aromatic products
soot
carbon recovery
20

0
800 850 900 950 1000

100

80
Yield ( % carbon input)

60

residual 2MN
40 aromatic products
soot
carbon recovery
20

0
800 850 900 950 1000
Temperature (deg C)

Figure 3-2 The overall temperature series of pyrolytic reactions of 1-methylnaphthalene (top panel)
and 2-methylnaphthalene (bottom panel). “1MN” represents 1-methylnaphthalene; “2MN”
represents 2-methylnaphthalene; “carbon recovery” represents the sum of residual reactant,
aromatic products and soot on carbon basis.

Figure 3-3 represents the pyrolytic product yields of the methylnaphthalenes. The C10H8 isomers,

with more than 98% as naphthalene, are the most abundant for all the temperatures. The other

C10H8 isomer identified is benzofulvene, which may be formed by naphthalene isomerization6.

Subsequent loss of CH2 from benzofulvene results in indene formation, which is also identified

in minor quantities. 2-Methylnapthalene is the second most abundant product at temperatures

below 950ºC, and the decrease at higher temperatures may be due to its increased consumption

as a reactant. Minor quantities of four C12H10 isomers and nine C12H12 isomers are also identified

37
and shown in Figure 3-4. The four C12H10 isomers identified are: 1- and 2-vinylnaphthalene,

acenaphthene and biphenyl, in decreasing order. The vinylnaphthalenes may be resultant from

ethylnaphthalene decomposition. The nine C12H12 isomers identified are: 2- and 1-

ethylnaphthalene, 2,6-, 1,7-, 1,6-, 1,4-, 1,2-, 1,5-, and 1,8-dimethylnaphthalene, of which 1,4-

and 1,2-dimethylnaphthalene have been identified by Platonov7 and 1,6-dimethylnaphthalene by

Mikolalczak8. The increase and then decrease trend with temperature is an indication of the

intermediate nature of these compounds. The formation of acenaphthalene (C12H8) may be due to

cyclodehydrogenation of the above C12 compounds. Up to 1.3% of three binaphthyl isomers are

identified from 1-methylnaphthalene pyrolysis, with the 2,2’ binaphthyl being the most abundant

and 1,1’ binaphthyl the least abundant.

100

10
Yield ( % carbon input )

0.1
C10H8(2)
2-MN
C12H10(4)
0.01 C12H12(9)
C12H8
C20H14(3)
0.001
800 850 900 950 1000

100

10
Yield ( % carbon input )

0.1
C10H8(2)
1MN
C12H10(4)
0.01 C12H12(7)
C12H8
C20H14(3)
0.001
800 850 900 950 1000
Temperature (deg C)

38
Figure 3-3 Yields of aromatic products in the pyrolysis of 1-methylnaphthalene (top panel) and 2-
methylnaphthalene (bottom panel). Numbers in parenthesis denote the number of isomers. “1MN”
represents 1-methylnaphthalene; “2-MN” represents 2-methylnaphthalene.

The product formation from 2-methylnaphthalene pyrolysis shown in the bottom panel of Figure

3-3 is analogous to those from 1-methylnaphthalene with the difference that the isomerization of

the reactant resulted in 1-methylnaphthalene formation as the second most abundant product.

The four C12H10 isomers identified are: 2- and 1-vinylnaphthalene, acenaphthene and biphenyl, in

decreasing order. The seven C12H12 isomers identified and shown in Figure 3-5 are: 2-

ethylnaphthalene, 2,6-, 1,7-, 1,6-, 1,4-, 1,5- and 1,2-dimethylnaphthalene in decreasing order.

The binaphthyl isomers are also identified from 2-methylnaphthalene pyrolysis, and are resultant

from naphthalene pyrolysis.6


1,7-DMN

8
1,6-DMN
kCounts

Ion: 156 C12H12


7
6
1,4-DMN
2-EN

1-EN

5
1,5-DMN
2,6-DMN

4
1,2-DMN

3
1,8-DMN

2
1
0

10.0
kCounts

Ion: 154 C12H10


2-Vinylnaphthalene
1-Vinylnaphthalene

7.5
Biphenyl

Acenaphthene

5.0

2.5

0.0

17.5 18.0 18.5 19.0 19.5


minutes

Figure 3-4 The selective ion chromatogram for C12H12 and C12H10 isomers in
1-methylnaphthalene pyrolysis. “EN” represents ethylnaphthalene and “DMN” represents
dimethylnaphthalene.

39
Ion: 156 C12H12(7)

kCounts
3.0

2EN
2.5

1,7-DMN
2,6-DMN
2.0

1,6-DMN
1.5

1,4-DMN
1.0

1,2-DMN
1,5-DMN
0.5

0.0

14 15 minutes

Figure 3-5 The selective ion chromatogram for C12H12 in 2-methylnaphthalene pyrolysis. “EN”
represents ethylnaphthalene and “DMN” represents dimethylnaphthalene.

3.3.2 The reactivity of 1- and 2-methylnaphthalene

The close resemblance of the reactant consumption and product formation patterns (Figure 3-2

and Figure 3-3) of the two methylnaphthalenes indicates that the same reaction mechanisms are

governing the pyrolysis of these two compounds. Figure 3-6 suggests 1-methylnaphthalene’s

slightly higher reactivity than its isomer, which is evidenced by the faster reactant consumption

and more product formation at the same temperatures. This is consistent with the Dewar activity

number: The smaller Dewar number of 1-methylnaphthalene (1.81; 2.12 for 2-methylnaphtalene)

indicates its higher reactivity than its isomer 9, 10.

40
100

residual reactant
products
80 isomer formed

Yield (% carbon input)


60

40

20

0
800 850 900 950 1000
Temperature (deg C)

Figure 3-6 The comparison of pyrolytic reactivity of the two methylnaphthalenes. The solid lines
and solid markers represent 1-methylnaphthalene data; the dashed lines and hollow markers
represent 2-methylnaphthalene data.

3.3.3 The hypothesized reaction mechanism of 1-methylnaphthalene pyrolysis

Figure 3-7 is a high level description of the hypothesized reaction mechanisms from 1-

methylnaphthalene pyrolysis. The pyrolytic experiments are initiated by the homolysis of the

reactant and the formation of the naphthylmethyl and H radical. The formation of naphthalene

can be resultant from demethylation, as well as the H displacement of the methyl group. The

latter becomes dominant with an increased H radical pool at elevated temperatures3. Naphthalene

can isomerize to form benzofulvene and then to indene by the loss of the CH2 radical6. The

dimethylnaphthalenes are resultant from methyl radical recombination to the methylnaphthalenes.

Isomerization results in 2-methylnaphthalene; it first increases with temperature and then

decreases, an indication of the decomposition of the other isomer at elevated temperatures. The

formation of 1-ethylnaphthalene (2-ethylnaphthalene for 2-methylnaphthalene) is probably due

to the methyl addition to the resonance-stabilized radical; the subsequent loss of hydrogen results

in 1-vinylnaphthalene (2-vinylnaphthalene for 2-methylnaphthalene). Formation of

41
acenaphthalene is due to the cyclodehydrogenation of acenaphthene, and it can also be formed

by two-carbon addition to the 1-naphthyl radical followed by cyclodehydrogenation3. The

product formation from 2-methylnaphthalene pyrolysis is analogous to that of 1-

methylnaphthalene.

CH3 CH2
CH2
+ CH3 -2H

1-ethylnaphthalene 1-vinylnaphthalene

-2H

acenaphthalene acenaphthene

CH3 + C 2 H2 - H
CH2
C +H

(isomerize)
naphthalene benzofulvene
(isomerize)
3 binaphyls

CH3

2-methylnaphthalene indene

Figure 3-7 Hypothesized pathways of 1-methylnaphthalene pyrolysis.

The formation of C10H8 isomers increases with temperature for pyrolysis experiments. The rate

of increase slows down at higher temperatures, which may be due to the fragmentation of these

isomers. The decreased recovery rates between 925 and 975°C may be due to the increased

formation of methyl radicals, which tend to form ethane through recombination2 and escaped the

current experimental setup. At temperatures higher than 950°C, soot formation increased rapidly,

42
while the formation of dimethylnaphthalenes and ethylnaphthalenes tended to decrease. Soot

increase might be due to naphthylmethyl radical addition/recombination, which improved

product recovery.

Isomerization first increases with temperature increase, to a certain point, and then decreases

with further temperature increase due to the decomposition of the isomer products. Observations

of the much-studied catalytic 1-methylnaphthalene isomerization in solutions have indicated that

the methylnaphthalene isomerization is initiated by proton addition to the 1-position followed by

dealkylation (intramolecular shift of the methyl group)11. Therefore, it is plausible to assume that

in radical assisted systems such as ours, isomerization is resultant from both H addition and

intramolecular alkyl shift. This assumption is supported by computational calculations (using

PM3) that the electron density on the 2-position of the 2-methylnaphthalene (the 1-position of

the 1-methylnaphthalene) is the highest among all of the carbon atoms, that is, the most

favorable position for H addition. However, in the oxidation of 2-methylnaphthalene, the

intramolecular alkyl shift may be inhibited by competitive oxidative pathways.

3.3.4 The formation of carbon doubling products

Unlike toluene pyrolysis, the pyrolysis of the methylnaphthalenes only resulted in minor

quantities of carbon dimerization products with 20-22 carbon atoms. Three C22 compounds

identified from the pyrolysis of methylnaphthalenes are: indeno(1,2,3-cd)pyrene, benzo(g,h,i)-

perylene, and dibenz(a,h)anthracene, with the latter two qualitatively agreeing with Lijinsky et

al5. PAH products higher than C22 were not found by HPLC analysis. Three binaphthyl (C20)

isomers have been identified as indicated earlier. The total yield of these products from 1-

43
methylnaphthalene pyrolysis is approximately 4.96% at 950ºC, and 1% for 2-methylnaphthalene,

which is in good agreement with the 1.53% from an earlier study on 2-methylnaphthalene

pyrolysis5. More carbon dimerization products from 1-methylnaphthalene are also an indication

of its higher reactivity over its isomer.

The lack of significant quantities of carbon doubling products may be due to the continued

growth of these products into soot. This is consistent with other studies and is also consistent

with the hypothesized formation mechanisms described earlier.

3.3.5 Soot particulate size distribution

The particulate size distribution of soot was measured using a personal cascade impactor in the

experiments of 1-methylnaphthalene pyrolysis at 950 and 1000°C, since soot became a major

product therein. At 950ºC, the products collected included both tar and soot, and only existed in

the last two substrates and the backup filter, as indicated in Table 3-1. At 1000 ºC tar did not

exist in the experiment and the sizes of most soot formed at 1000°C also are smaller than 1 μm,

even though the detailed distribution is different from that at 950°C. It is in good agreement with

the study on particles formed in polystyrene combustion by Durlak et al12, where most of

particles are smaller than 1 μm. The distribution of soot size became more dispersive when

temperature was increased to 1000°C. At higher temperature more soot are formed and their

growing-up change become larger through collision and coagulation.

Table 3-1 Distribution of soot size formed in 1-methylnaphthalene pyrolysis at 950 and 1000°C .

Stage Cut-point Percent in Size Range (%)


Dp (um) 950 1000

44
5 3.5 0 2.99
6 1.55 0 16.41
7 0.93 41.77 33.58
8 0.52 53.05 28.73
Backup 0 6.1 18.3
Filter

3.3.6 Detection of gaseous products

The final emission gases of pyrolysis of 1-methylnaphthalene were collected using SKC sample

bags and analyzed by our GC/MS. Only a very small amount of benzene vapor, which accounted

for less than 0.3% carbon input, was detected. Methane was found in the emission gas using

another GC system (HP 5890 series II equipped with a dual column HP 10 ft BX-45/60 mesh

and HP 6 ft HAYESEP Q 80/100 mesh, and TCD detector), but its portion of carbon input was

less than 0.5%. In order to simplify tasks, the products escaping with the final emission gas are

neglected and emission gas will not be collected in most of the pyrolysis experiments.

3.4 Conclusions

The formation of resonance-stabilized naphthylmethyl radicals initiated the PAH and soot

formation mechanisms of the methylnaphthalenes, and it provided the hydrogen radical pool for

the formation of naphthalene and methyl loss. The methyl radical recombines with the reactants

to form dimethylnaphthalenes, ethylnaphthalenes, and subsequent products. The lack of carbon

dimerization products in significant quantities may be due to their continued growth into soot.

The reactions by the methyl group have added complexities to the thermal decomposition of the

methylnaphthalenes in comparison to other resonance-stabilized compounds, such as toluene and

45
indene. The close resemblance of the reactant consumption and product formation of the

methylnaphthalenes indicates that the same reaction mechanisms are governing the pyrolysis of

these compounds. 1-Methylnaphthalene exhibits slightly higher reactivity than 2-

methylnaphthalene.

References

1. Shaddix, C. R.; Brezinsky, K.; Glassman, I. In Oxidation of 1-Methylnaphthalene, 24th


Symp. Int. Combust., 1992; The Combustion Institute: 1992; pp 683-690.

2. Shaddix, C. R.; Brezinsky, K.; Glassman, I., Analysis of Fuel Decay Routes in the High-
Temperature Oxidation of 1-Methylnaphthalene. Combust. Flame 1997, 108, 139-157.

3. Pitsch, H., Detailed Kinetic Reaction Mechanism for Ignition and Oxidation of alfa-
Methylnaphthalene. In 26th Symp. Int. Combust., The Combustion Institute: 1996; pp 721-728.

4. Goos, E.; Douce, F.; Djebaili-Chaumeix, N.; Braun-Unkhoff, M.; Slavinskaya, N.; Frank,
P.; Paillard, C. In Experimental and numerical investigation focusing on early soot formation
of toluene and 1-methylnaphthalene at high pressures, Proceedings of the European Combustion
Meeting, 2003; 2003.

5. Lijinsky, W.; Raha, C. R., The Pyrolysis of 2-Methylnaphthalene. J. Org. Chem 1961, 26,
3566-3568.

6. Lu, M.; Mulholland, J. A., PAH Growth from the Pyrolysis of CPD, Indene and
Naphthalene Mixture. Chemosphere 2004, 55, 605-610.

7. Platonov, V. V.; Ivleva, L. N.; Vol-Epshtein, A. B.; Kasimtseva, T. V.; Klyavina, O. A.,
Kinetics of the Homogenous Pyrolysis of Naphthalene and Monomethylnaphthalenes. Solid Fuel
Chemistry (English Translation of Khimiya Tverdogo Topliva) 1986, 20, 74-82.

8. Mikolajczak, C. J.; Wornat, M. J.; Dryer, F. L.; Fringer, O. B.; Held, T. J., Chemical and
Physical Processes in Combustion 1996, 183-186.

9. Smith, C. M.; Savage, P. E., Reactions Of Polycyclic Alkylaromatics - Structure And


Reactivity. Aiche Journal 1991, 37, (11), 1613-1624.

10. Dewar, M. J. S., A Molecular Orbital Theory of Organic Chemistry. VI. Aromatic
Substitution and Addition. J. Am. Chem. Soc. 1952, 74, (13), 3357-3363.

46
11. Ferino, L.; Monaci, R.; Rombi, E.; Solinas, V., Microcalorimetric investigation of
mordenite and Y zeolites for 1-methylnaphthalene isomerisation. J. Chem. Soc., Faraday Trans.
1998, 2647-2652.

12. DurLak, S. K.; Biswas, P.; Shi, J.; Bernhard, M. J., Characterization of polycyclic
aromatic hydrocarbon particulate and gaseous emissions from polystyrene combustion. environ.
Sci. Technol. 1998, 32, 2301-2307.

47
4 CHAPTER IV 2-METHYLNAPHTHALENE OXIDATION

4.1 Introduction

In Chapter III, pyrolysis of 1- and 2-methylnaphthalene was studied experimentally. Aromatic

products were collected by an ice bath trap system with DCM, and identified by GC/MS. Soot

was either collected by filter or scraped from the inner surface of the quartz tubing reactor, and

then weighed. The product yields as the function of temperature were demonstrated and the

pathway of 1-methylnaphthalene pyrolysis was hypothesized. However, pure pyrolysis condition

is not practical in most of combustion. Due to the attack of oxygen and hydroxide radicals which

are more active than radicals formed in pure pyrolysis, fuel decomposition and product formation

happen at lower temperatures in oxidation condition. In this chapter, 2-methylnaphthalene

oxidation is studied experimentally. The experimental condition is the same as that for the

pyrolysis experiments except that the temperature range is between 650 and 950°C with 50°C

increments and the carrier gas is 3470 ppm oxygen in nitrogen. The average equivalence ratio is

2 (fuel rich) so that the effects of both pyrolysis and oxidation can be observed. The input

reactant concentration is 472-611 ppm. CO concentration is detected by an emission analyzer

(Testo 350).

48
4.2 Results and Discussion

4.2.1 Products and their yields as functions of temperature

One Chromatogram of products from 2-methylnaphthalene oxidation at 900ºC is shown in

Figure 4-1 and parts of products are labeled. The experimental data of product yields are shown

in Table B-3 for 2-methylnaphthalene oxidation. The reaction profiles of 2-methylanpthalene

oxidation are shown in Figure 4-2. Up to 35.1% of aromatic products and 2.9% of CO are

formed at 950ºC. The recovery rates range from 90.1% at 650ºC to 38.0% at 950ºC, due to the

increased oxidation reactions at higher temperatures. The yields of major aromatic products in

the 2-methylnaphthalene oxidation are shown in Figure 4-3: products that are similar to pyrolysis

are plotted in the top panel, and decomposition products are plotted in the bottom panel. The

major products identified include naphthalene, 2-ethylnaphthalene, 2-naphthylaldehyde, toluene,

indene, and phenylacetylene. The products in the bottom panel generally increase with

temperature, an indication of increased oxidation reactions which favors the decomposition of

the reactants and intermediates into smaller sizes. The six C12H12 isomers identified and shown in

Figure 4-4 are: 2-ethylnaphthalene, 2,6-, 1,7-, 1,6-, 1,4- and 1,2-dimethylnaphthalene, with 2-

ethylnaphthalene being the most abundant.

49
9 11 12
RIC all 2m900_3_30002.SMS
200
kCounts

5
150
13

100

50
1 16
6
15
10
14 17
19
18
4 7 8 20
3
0

5 10 15 20
minutes

Figure 4-1 The chromatogram of products in 2-methylnaphthalenes oxidation at 900°C. 1 benzene;


2 toluene; 3 ethylbenzene; 4 p-xylene; 5 phenylacetylene; 6 styrene; 7 benzaldehyde; 8
benzofuran; 9 indene; 10 benzofulvene; 11 naphthalene; 12 2-methylnaphthalenes; 13 1-
methylnaphthalenes; 14 C12H10s & C12H12s; 15 acenaphthalene; 16 2-naphthaldehyde; 17
fluorene; 18 C13H10s; 19 phenanthrene; 20 anthracene

100

80
Yield (% carbon input)

60
aromatic products
40 CO
residual 2M
carbon recovery
20

0
650 700 750 800 850 900 950
Temperature (deg C)

Figure 4-2 The overall temperature series of 2-methylnaphthalene oxidation. “2M” represents 2-
methylnaphthalene.

50
100
C12H12(6)
1MN
C10H8(2)
Yield (% of carbon input)
10 C11H8O
C12H10(4)

0.1

0.01
650 700 750 800 850 900 950
Te mpe rature (deg C)

100
Benzene
Toluene
Phenylacetylene
Yield (% of carbon input)

10 Stylene
Indene

0.1

0.01
650 700 750 800 850 900 950
Temperature (deg C)

Figure 4-3 Yields of aromatic products in 2-methylnaphthalene oxidation. Numbers in parenthesis


denote the number of isomers. “1MN” represents 1-methylnaphthalene.

51
kCounts
Ion: 156 C12H12(6)
2.5

2-EN
2.0

1,7-DMN
1.5

2,6-DMN

1,6-DMN

1,4-DMN
1.0

1,2-DMN
0.5

0.0

14.25 14.50 14.75 15.00 15.25 15.50 15.75


minutes

Figure 4-4 The selective ion chromatogram for C12H12 in 2-methylnaphthalenes oxidation. “EN”
represents ethylnaphthalene and “DMN” represents dimethylnaphthalene.

4.2.2 The hypothesized reaction mechanisms

Compared with the pyrolysis of 2-methylnaphthalene at the same temperature range (800-950ºC),

higher reactant consumption rates are observed in oxidation. This can be accounted for by the

presence of the oxidative species, such as O, OH, and oxygen, which is also consistent with the

oxidative studies on indene66 and 1-methylnaphthalene57, 58. The presence of oxygenated species

accelerated the hydrogen abstraction from the reactant, and the formation of naphthylmethyl

radicals up to 800ºC, which is evidenced by the increase of C12H12 isomers. The majority of

C11H8Os is 2-naphthaldehyde, with minor quantities of 1-naphthaldehyde at 900 and 950ºC. The

similar temperature series of C12H12s and that of C11H8Os indicates similar formation routes.

The product formation in 2-methylnaphthalene oxidation follows the pathways of methyl loss,

methyl addition, and the oxidation of naphthylmethyl, which results in two C10H8 isomers, 2-

ethylnaphthalene and 2-naphthaldehyde, respectively. Indene is mainly formed from the


52
oxidation of naphthalene followed by CO elimination from the naphthoxy radical57, 58
, in

addition to naphthalene decomposition observed in 2-methylnaphthalene pyrolysis 67.

The isomerization of the reactant (2-methylnaphthalene) to 1-methylnaphthalene is not as

apparent as in the pyrolytic experiments. The almost constant yields over the temperature range

suggest that it is neither a reactant nor a product. The reactant purity analysis suggests that 1% 1-

methylnaphthalene is the impurity of 2-methylnaphthalene.

CH3 CH3
CH2
+ CH3

2-ethylnaphthalene
- CH3

C - H - CO O

2-naphthaldehyde
+H

CH2
-CH3

(isomerize)

naphthalene benzofulvene indene

Figure 4-5 Hypothesized pathways of 2-methylnaphthalene oxidation

Reference
1. Lu, M. Carbon growth from cyclopentadienyl moities in combustion. Georgia Institute of
Technology, Atlanta, GA, 2000.

2. Shaddix, C. R.; Brezinsky, K.; Glassman, I., Oxidation of 1-Methylnaphthalene. In 24th


Symp. Int. Combust., The Combustion Institute: 1992; pp 683-690.

53
3. Shaddix, C. R.; Brezinsky, K.; Glassman, I., Analysis of Fuel Decay Routes in the High-
Temperature Oxidation of 1-Methylnaphthalene. Combust. Flame 1997, 108, 139-157.

4. Lu, M.; Mulholland, J. A., PAH Growth from the Pyrolysis of CPD, Indene and
Naphthalene Mixture. Chemosphere 2004, 55, 605-610.

54
5 CHAPTER V GLOBAL KINETIC ANALYSIS OF
METHYLNAPHTHALENE PYROLYSIS

5.1 Introduction

While it is essential to understand the formation pathways of soot and PAHs, the combustion of

fuels, even a simple compound, is a complex process involving a series of sequential and parallel

reactions. Even though a detailed model including relevant species and the reactions between

them could be constructed to describe the process accurately1, 2, the complexity of the method

and the need for prior knowledge of detailed reaction mechanisms offset its advantage. In

addition, the determination of the kinetic parameters of elementary reactions is challenging.

Comparatively, global kinetic analysis, which considers only the overall reaction including a

single or a small number of reactions, can provide a simpler alternative to understand the

reactivity of the target compounds 3, 4.

In this chapter, the pyrolysis of the methylnaphthalenes has been quantitatively studied using the

global kinetic method to further understand the process. The used data come from our previous

experimental study, where the pyrolysis experiments of the methylnaphthalenes were performed

ranging from 800 to 1000ºC, with the temperature series of the major products provided and

product formation pathways hypothesized5. The equations for reactant consumption and product

formation were derived. The kinetic parameters of methylnaphthalene decomposition and


55
isomerization were obtained as the isomerization process which occurred in pyrolysis gave rise

to the second abundant products.

Tabular flow reactor, which was employed in the experiments, has been used extensively in

combustion chemistry and many of them operate in the laminar flow region. The simplest

mathematic description for tabular flow reactors is plug flow reactor (PFR), where the flow is

one-dimensional with no radial variations in concentration and velocity6. Turbulent flow without

back mixing is a good example of plug flow. For laminar flow, the plug flow assumption will not

be applicable if the radial diffusion of species is neglected, considering the parabolic velocity

profile for a fully developed laminar flow. However, it becomes valid if the radial diffusion

approaches infinity and the reactor become a plug flow reactor. PFR assumption for laminar flow

experiments has already been adopted in kinetic studies of thermal decomposition of different

organic compounds, as listed in Table 5-1, and good results were demonstrated.

Table 5-1. Kinetic studies on pyrolysis of organic compounds based on plug flow assumption.

Author Reactant Experiment details Results


Reactor: L: 0.295 m, i.d. 8 mm
Reactant decomposition
Carrier gas: 8.5% H2 in He
7 Tert- lnA Ea
Cutler et al. Reynolds number : 16 – 40
Butylalcohol (kCa/mol)
Temperature: 700-810 °C
27.7 53.3 ± 11
Retention time: 0.07-0.15 s
Product formation
Product logA Ea
Reactor: L: 1.22 m, i.d. 2 mm
(kCa/mol)
Carrier gas: N2
Ledesma et Benzene 10.9 54.1
Catechol Reynolds number : 33-69
al4 Naphthalene 14.7 74.3
Temperature: 500-1000 °C
Phenanthrene 16.4 84.3
Retention time: 0.4 s
Anthracene 18.1 92.9
Pyrene 16.4 86.6

56
2- 16.6 83.9
vinlynaphthalene
2- 14.8 77.8
ethylnaphthalene
Reactant decomposition
Reactor: L: 1.5 m, i.d. 52.5 mm
Carrier gas: N2
Kershenbaum lnA Ea (kJ/mol)
n-butane Reynolds number : 80-160
and Leaney8 24.9 213 ± 14
Temperature: 640-7400 °C
24.5 212 ± 12
23.4 197 ± 19

On the other hand, if the axial diffusion reaches infinity, a laminar flow will become well back-

mixed and the reactor will become a continuous-stirred tank reactor (CSTR). Therefore the flow

characteristics should be evaluated carefully before modeling it. In this study, a series of criteria

summarized by Cutler and his colleagues 7 are used to demonstrate that the tubular reactor used

in our experiments can be idealized as a plug flow reactor with fast radial diffusion and

insignificant axial diffusion. The result deviation caused by the idealization is discussed and

proved to be in the acceptable range.

First order reaction law has been used for description of reactant conversion in thermal

decomposition of 9-methylphenanthrene9 and 1-methylpyrene10, which are the

methylnaphthalenes’ homolog. Furthermore, it was shown that first order reaction provided a

good representation of the experimental data of reactant decomposition in the study of the 1-

methylnaphthalene pyrolysis in a batch reactor in the temperature range from 380 to 450 ºC with

several tens of hours retention time11. Therefore, we use first order rate law in this study.

57
When reaction temperature is lower than 850 ºC, there were no soot or semi-volatile products

observed on the inner surface of the reactor. At higher temperatures, soot was formed and

deposited on the tube wall. It is possible that wall reactions happen also: some semi-volatile

products are formed directly from reactant decomposition and attached to soot. It can be checked

by analyzing the chemical composition of the rinse DCM solution washing the reactor inner

surface after pyrolysis. Unfortunately, this step wasn’t conducted in our experiments. In this

study we don’t consider possible wall reactions or heterogenic reactions, which means that our

modeling results should be used with reservations until it is proved that wall reactions are really

negligible.

5.2 Validation of the Plug Flow Assumption

Cutler and his colleagues7 summarized the others’ works which overcame the limitations

accompanying the use of the plug-flow treatment of tubular-flow reactor and assured the

legitimacy of the plug-flow idealization. Those criteria cited from Cutler’s paper are listed in the

Table 5-2. The values of non-dimensional number were used to ensure that the difference

between the species concentrations or rate constant in plug flow model and their true values is

only several percent. The former four are employed to determine whether axial diffusion can be

neglected, and the latter four are for checking whether the radial diffusion can “smear” the

parabolic velocity profile.

58
7
Table 5-2. Criteria for the validity of the plug flow idealization summarized by Cutler et al.

NO. Non-dimensional number Criteria Sources


1 τ fc ,R 2 / (τ ckτ sd ,R ) Dickens et al., 1960
<0.1 Howard, 1979
Furue and Pacey, 1980
2 τ fc, L / τ sd , L <0.06
Mulcahy and Pethard,
1963
3 τ fc, R / τ sd , R <0.02
Dang and Steinberg,
1980
4 τ sd , R /(48 τ ck ) + τ fc ,R 2 / (τ ckτ sd ,R ) <<1 Furue and Pacey, 1980
5 τ sd , R / τ fc, R <100
Walker, 1961
Brown, 1978
6 τ sd , R / τ fc, L <0.5
Cleland and Wilhelm,
1956
7 τ sd , R / τ ck Walker, 1961
Vignes and Trambouze,
1962
<1
Ogren, 1975
Lede and Villermaux,
1977
8 τ fc , R / τ ck <0.05 Brown, 1978

Those characteristic times and relative notations are defined as following:

τ fc , R = R / V

τ fc , L = L / V

τ sd ,R = R 2 / D

τ sd , L = L2 / G

τ ck = k −1

D: the diffusion coefficient of the chemical species, m2/s

G: axial effective diffusion coefficient, G= D + R 2 L2 /(48 Dt 2 ) 12, m2/s

k: reaction rate constant, 1/s

59
L: reactor length, m

~
M : molecular weight

R: reactor radius, m

V : mean velocity, m/s

σ : molecular property characteristic of Chapman-Enskog theory

τ : characteristic time, s

t: retention time, s

p: gas pressure, atm

Subscripts

fc: forced convection

sd: species diffusion

ck: chemical kinetic

R, and L: radial or axial direction

Reynolds number in our experimental setup differed from 7 to 14 in the range of 600 to 1000°C.

It proves that the flow was in laminar status. In order to check the validation of PFR, those non-

dimensional numbers at different temperatures are calculated and compared with those criteria.

In this study, D is the diffusion coefficient of methylnaphthalenes in the carrier gas helium. It

was estimated by using the Chapman-Enskog theory13 and the Lennard-Jones potential

parameters of naphthalene14 were used due to the unavailability of methylnaphthalene data.

Helium viscosity were obtained from Yaws’ figures15.


60
Here is one example of calculation for 1-methylnaphthalene pyrolysis experiment at 1000ºC. The

retention time is one second and some known parameters are:

L = 0.61 m; R = 0.0085 m; γ = 494 micropoise ; V =0.61 m/s; T= 1273 K; p= 1 atm; t= 1s

τ fc , R = R / V = 0.0085/.61 = 0.0156 s; τ fc , L = L / V = 1 s; τ sd , R = R 2 / D

The diffusion coefficient of the reactant (1-methylnaphthalene, C11H10) in helium was estimated

using the Chapman-Enskog equation.13


~ ~
0.00186T 3 / 2 (1 / M 1 + 1 / M 2 )1 / 2
D=
pσ 122 Ω

~ ~
M 1 = 2, M 2 = 142

Use naphthalene data14 for 1-methylnaphthalene, σ =6.45, (ε / k B ) Naphthalene = 554.4

data in Table 5.1-2 13 for helium, σ =2.551, (ε / k B ) He =10.22

Then σ 12 =( σ 1 + σ 2 )/2 = (6.45+2.55)/2 = 4.5

ε 12 / k B T = (ε 1 / k B ) Naphthalene (ε 2 / k B ) He / T = 554.4 × 10.22 / 1273 =0.0591

From Table 5.1-313, Ω =0.688

0.00186 × 12733 / 2 (1 / 4 + 1 / 142)1 / 2


D= = 3.07 cm2/s
4.5 × 0.688
2

τ sd , R = R 2 / D = 0.0085 2 / .000307 = 0.24s

G = D + R 2 L2 /(48 Dt 2 ) = 0.000307+0.00852*0.612/(48*.000307) = 0.00213

τ sd , L = L2 / G = 0.612 / 0.00173 = 174.7 s

61
We will use our kinetic analysis results based on the assumption that the idealization of the plug

flow reactor is feasible, which is being proven in this chapter.

k = Aexp(-Ea/RT) = exp(22.92-224/(0.008314*1273)) = 5.78 s-1

τ ck = k −1 = 0.199 s

Based on the above data and method, non-dimensional numbers are calculated from 600 to

1000ºC and shown in Table 5-3 with some flow properties. It was discovered that at temperature

lower than 1000 ºC almost all the criteria for the validity of plug flow idealization have been

satisfied except τ fc, R / τ sd , R . We will show that the error of kinetic parameters caused by the

unmet criteria is less than 12% in the latter discussion section, and the solution with PFR

assumption is much simpler than that with back mixing consideration. For two criteria τ sd , R / τ fc , R

and τ sd , R / τ fc , L where values are close to threshold, the estimated errors for reaction rate constant

is less than 5%16, 17


and 7%18 separately. Error for activation energy will be less than 0.5%

considering logarithm relationship between it and reaction rate constant. Therefore, the reactor is

treated as a plug flow reactor in this study.

Table 5-3. Characteristic times and non-dimensional numbers of the experimental setup from 600
to 1000°C.

T(°C) 600 700 800 900 1000


ε/(kT) 0.086 0.077 0.070 0.064 0.059
kT/ε 11.6 12.9 14.3 15.6 16.9
Ω 0.73 0.72 0.71 0.70 0.69
D (m2/s) 1.65E-04 1.96E-04 2.31E-04 2.68E-04 3.07E-04
G (m2/s) 3.57E-03 3.05E-03 2.66E-03 2.36E-03 2.13E-03
τ fc , R = R / v (s) 0.0156 0.0156 0.0156 0.0156 0.0156

62
τ fc , L = L / v (s) 1 1 1 1 1
τ sd ,R = R 2 / D (s) 0.44 0.37 0.31 0.27 0.24
τ sd , L = L2 / G (s) 104.3 122.1 140.0 157.7 174.7
τ ck = k −1 (s) 2812.92 117.94 8.93 1.05 0.17

τ fc ,R / (τ ckτ sd , R )
2
<0.1 1.57E-07 4.48E-06 6.95E-05 6.85E-04 4.78E-03
τ fc , L / τ sd , L <0.06 0.0096 0.0082 0.0071 0.0063 0.0057
τ fc , R / τ sd , R <0.02 0.032 0.038 0.045 0.052 0.059

τ sd , R /(48 τ ck ) + τ fc , R / (τ ckτ sd , R )
2
<<1 3.41E-06 6.94E-05 8.00E-04 6.04E-03 3.31E-02
τ sd , R / τ fc , R <100 31.5 26.4 22.5 19.4 16.9
τ sd , R / τ fc , L <0.5 0.44 0.37 0.31 0.27 0.24
τ sd , R / τ ck <1 1.56E-04 3.12E-03 3.50E-02 2.57E-01 1.36E+00
τ fc , R / τ ck <0.05 4.95E-06 1.18E-04 1.56E-03 1.33E-02 8.06E-02

If the retention time increases to 100 second, which means a slower flow, the first four and the

last non-dimensional numbers increase greatly to cross the threshold. At that situation, axial

diffusion has to be considered in modeling. In contrast, if the flow average velocity increases 10

τ sd , R τ fc , R τ sd , R τ ck
times and retention time becomes 0.1 second, / and / become larger and their

criteria aren’t met. It is because the flow is so fast that the radial diffusion couldn’t smear the

parabolic velocity profile. Certainly, if the flow rate is large enough to make it become turbulent

flow, plug flow assumption will be valid if axial diffusion is proved negligible.

5.3 Basic Models of Reactant Consumption and Product Formation

Global kinetic analysis can be useful in providing a general understanding of the combustion

process of practical fuels without the detailed chemical analysis of the end products. Using this

method, it is possible to predict the emission of hydrocarbon pollutants in the combustion of

63
practical fuels once the characteristics of formation and decomposition of those pollutants are

known and optimization work can be processed to minimize pollutant emission. The simplest

model that describes pyrolytic reactions is a first order reaction4, 19, which is suitable for single

reactant systems in most circumstances. In fact, first order rate law was proved applicable for the

pyrolysis of 1-methylnaphthalene at around 400 °C.11 Therefore, it is feasible to assume that the

pyrolysis of methylnaphthalenes is a first-order reaction in this study.

For product formation,


dX
= k(X f − X ) (1)
dt
where X is the yield of products at time t. Xf is the ultimate yield when t → ∞ . All the yields are

expressed as fractions of the initial reactant input. k is the first-order reaction rate constant which

can be expressed by the Arrhenius equation:


Ea

k = Ae RT
(2)

where A is the pre-exponential factor (s-1), and Ea is the activation energy (kJ·mol-1). R is the

universal gas constant, and T is the temperature (K).

Substituting k with equation (2), the final integrated form of the rate expression (1) is shown as

equation (3) for product formation:


E
− a
− Ae
X = X f (1 − e
RT t
) (3)

For reactant decomposition,

64
dX
= −kX (4)
dt
where X is the fraction of the residual/unreacted reactant. The result of integration from 0 to time

t is:
E
− a

X = X 0 e − Ae
RT t
(5)

X0, the initial fraction of the reactant, is assumed as 100%.

Equation (3) and equation (5) can be linearized by taking the natural logarithm on the equation

twice with a little transformation to equation (6) and equation (7).

Xf Ea
ln(ln ) = ln At − (6)
Xf −X RT

X0 E
ln(ln ) = ln At − a (7)
X RT

A and Ea can be obtained in both linear and nonlinear approaches.

5.4 Results and Discussion

5.4.1 Reactant Decay in Pyrolysis of Methylnaphthalenes

With equation (7) and Microsoft Excel, the kinetic parameters for reactant consumption in one

reaction can be calculated based on experimental data. The results are shown in Table 5-4 under

the column “Linear”. The linearized model is generally a good method for estimation

considering its simple expression. It has been used to process the data of PAH formation from

catechol pyrolysis4 and good agreement was shown20.


65
With the development of more effective computational packages such as MatLab, the solution of

the nonlinear model also becomes simple. Fitting the experimental data to equation (5), the

kinetic parameters of methylnaphthalene consumption were obtained and listed in Table 5-4

under “Nonlinear”. The error estimation for model prediction and estimated model parameters

were processed and expressed by 95% confidence interval half-widths, as shown in Table 5-4.

The estimated kinetic parameters by the linear model are close to those by the nonlinear model.

The errors for 2-methylnapthalene are larger than those for 1-methylnaphthalene because 2-

methylnaphthalene experimental points are more dispersive than those of its isomer. Our results

of 1-methylnaphthalene is very close to that of Leininger and his colleagues 11, lnA = 22.8 and

Ea= 198 kJ/mol.

Table 5-4 Kinetic Parameters for reactant decay in pyrolysis of methylnaphthalenes

Reactant Nonlinear Linear

Ea Ea

lnA (kJ/mol) lnA (kJ/mol)

1-methylnaphthalene 22.9±3.4 224.09±32.70 22.03 215

2-methylnaphthalene 26.6±7.7 262.57±76.26 28.84 285

The results from the different methods with experimental data are shown in Figure 5-1. The

model prediction, no matter linear model or nonlinear model, is in good agreement with
66
experimental data. The sum of squared error (SSE) using the nonlinear method is 506.3 for 1-

methylnaphthalene decomposition and 343.1 for 2-methylnaphthalene decomposition.

Comparing with 527.5 for 1-methylnaphthalene and 375.4 for 2-methylnaphthalene using the

linear model, we have shown that the nonlinear method gives better fit over experimental data

although it is very close to the linear one. Moreover, it has already been pointed out that

linearization can bias parameter estimates since the equation transformation can distort the error

structure of the original measurements21, 22. Therefore, the nonlinear model is recommended for

more reliable results even though methods of adjustments for the linear model were provided to

better fit the experimental data23.

67
100
1-Methylnaphthalene

Yield (% carbon input)


80

60

40

20

0
800 850 900 950 1000

100
Yield (% carbon input)

80 2-Methylnaphthalene

60

40

20

0
800 850 900 950 1000
Temperature (deg C)

Figure 5-1 Comparison between experimental data (solid diamonds) and model prediction for
reactant consumption in 1-methylnaphthalene pyrolysis (top panel) and 2-methylnaphthalene
(bottom panel). The thick solid lines are result of the nonlinear model and dashed lines give their
95% confidence interval half-widths. The thin solid lines are result of the linear model.

5.4.2 The influence of axial diffusion on the kinetic analysis

The criterion which isn’t met was given by Dang and Steinberg17 to evaluate the effect of axial

diffusion. They also gave the mathematic description of first order homogeneous reaction

happening in a laminar flow tubular reactor for an isothermal system

68
r ∂C 1 ∂ ∂C ∂ 2C
2V [1 − ( ) 2 ] = D[ (r ) + 2 ] − kC ,
R ∂x r ∂r ∂r ∂x

And the boundary conditions of the system are

C (0, r ) = C 0

∂C
( x,0) = 0
∂r
∂C
−D ( x, R ) = k w C ( x, R )
∂r
Here C is the reactant concentration at the position (x, r) which are axial and radial coordinates,

and kw is the surface wall reaction rate constant. It is a partial differential equation and the given

solution contained both analytical and numerical parts, which made it very complex. Fortunately,

in his subsequent study, a dispersion model without consideration of radial diffusion was

provided to approximate the system. 24

d 2C dC
K2 2
+ K1 + ( K w − K )C = 0
dx dx

And the inlet and exit boundary conditions are


dC
K 1C + K 2 = K 1C 0 when x = 0 and C = 0 when x → +∞
dx
In this case, K w = 0 , K 1 = −V , K 2 = G , and K = kR 2 / D . The final analytical solution is

K2
−1+ 1− 4 (k w − k )
C 2 K 12
= exp[ x] (8)
C0 4K 2 K 2 / K1
1 + 1 − 22 (k w − k )
K1
This one dimensional dispersion model can describe the realistic phenomena in a limner flow
when K≤1 and kw≤D/R where kw is the wall reaction rate constant. These requirements are met in
this study. Fitting our experimental data with this equation, another set of kinetic parameters are
obtained. Ea is 241±33 and 280±77 kJ/mol for 1-methylnaphthalene and 2-methylnaphthalene

69
respectively. The relative difference between them and listed values in Table 5-4 is less than
10%. The obtained lnA is 26 for 1-methylnaphthalene and 26.6 for 2-methylnaphthalene, which
cause the relative difference with data list in Table 5-4 is less than 12%. Considering the
complex expression of equation (8), the equation based on PFR assumption will be employed
even though theoretically equation (8) is able to describe real scenario more accurately.

5.4.3 Hydrogen Concentration Estimate in 1-Methylnaphthalene Pyrolysis

There are four potential decomposition pathways for 1-methylnaphthalene pyrolysis, which are

listed in Table 5-5 along with their kinetic parameters cited from the literature25, 26
. The

estimated activation energy of 1-methylnaphthalene decomposition, 224 kJ/mol, is closer to

those of bond cleavage reactions, 343 and 406 kJ/mol, while it is one order of magnitude higher

than those of hydrogen abstraction, 32.77 and 21.55 kJ/mol. This may be an indication that the

first order assumption is more suitable for the simulation of bond cleavage rather than hydrogen

abstraction reactions.

Table 5-5 Possible reactions in 1-methylnaphthalene decomposition and their kinetic parameters. N
denotes the 1-naphthyl radical.

No. Reaction A Ea (kJ/mol) Reference


25
1 NCH3 + H Æ NCH2 + H2 1.26E14 32.77
25
2 NCH3 + H Æ NH + CH3 1.20E13 21.55
25
3 NCH3 Æ NCH2 + H 1.00E15 343
26
4 NCH3 Æ N + CH3 1.00E16 406

70
It is noteworthy that in the H/O system, the radical pool was in an approximate steady-state at

least in the first 50 ms after combustion starts.25 In that steady period, hydrogen concentration in

the experiments of 1-methylnaphthalene pyrolysis can be estimated with calculated reaction rates

and the assumption that the pyrolysis happens in steady state. For reactant:
dX
= −kG X = −k1 X [ H ] − k 2 X [ H ] − k 3 X − k 4 X
dt
kG − k3 − k 4
[H ] =
k1 + k 2

Here the subscript “G” denotes global kinetic analysis, and “1” is for reaction 1, “2” for reaction

2, “3” for reaction 3, and “4” for reaction 4. The above reaction rate constants can be calculated

with kinetic parameters obtained through kinetic analysis and listed in Table 5-4. The result

differs from 2E-14 to 7E-14 when the temperature is lower than or equal to 950ºC. The values

are much smaller compared, to 0.03-0.06 ppm, the values in the experiments of 1-

methylnaphthalene oxidation with 0.6-1.5 equivalence ratios.25 It is in agreement with Shaddix’s

suggestion that the hydrogen mole fraction decreases with increasingly fuel-rich conditions. The

calculation procedure also shows that homolysis reactions (3) and (4) became more important

and dominant when the temperature increases.

With respect to two homolysis reactions, the formation of naphthylmethyl radical should be the

dominant mechanism because of its low activation energy and resonance-stabilized properties.5

The comparison of the reaction rate constants of Reaction 1 through Reaction 4 calculated with

the provided kinetic parameters is shown in Figure 5-2. Reaction 1 is the major route of reactant

71
decomposition at temperatures lower than about 875°C and Reaction 3 become preponderant at

higher temperatures. This fact means that the homolysis reaction becomes more important at

higher temperatures. The reaction rate of Reaction 3 is one to two orders of magnitude larger

than that of Reaction 4 in the experimental temperature range. This also supports the conclusion

that formation of naphthylmethyl radical is the major pathway in pyrolysis before hydrogen

concentration has reached a certain level.5, 25

10
Reaction rate constant (s-1)

0.1

0.01
R1
R2
0.001
R3
R4
0.0001
800 850 900 950 1000
Temperature ( deg C)

Figure 5-2 Comparison of calculated reaction rate constant of reaction 1 through 4 listed in Table
5-5. “R1” represents reaction 1, “R2” denotes reaction 2, and so on.

5.4.4 The Yield of Intermediates in Pyrolysis

Many products tend to act as intermediates in both pyrolysis as well as oxidation; their yields

first increase with the increase of reactant decomposition and then decrease due to their own

decomposition when the temperature continues to rise. Equation (3) then becomes insufficient

for the intermediates as only product formation is calculated without the consideration of further

72
product decomposition. In our study, models of reactant decomposition and product formation

were combined to describe the yields of intermediates assuming that the process only involved

product formation and their further decomposition. In this study, the isomerization process (e.g.

2-methylnaphthalene formation in 1-methylnaphthalene pyrolysis) was used as an example, as it

resulted in the second most abundant product. Assuming that there is a certain proportion of 1-

*
methylnaphthalene which will be entirely converted to its isomer and its concentration is X 1M .

Then the series reaction is demonstrated as Figure 5-3.

1-methylnaphthalene ( X 1M
*
)

k1
2-methylnaphthalene (X2M)
k2
Pyrolysis products

Figure 5-3 The reaction model for 2-methylnaphthalene formation in 1-methylnaphthalene


pyrolysis. X1M denotes the concentration of 1-methylnaphthalene and X2M denotes the yield of 2-
methylnaphthalene.

The reaction series is expressed as the following

dX 1*M
= − k1 X 1*M (8)
dt

dX 2 M
= k1 X 1*M − k 2 X 2 M (9)
dt

The initial condition is X 2 M = 0 , X 1*M = X 1*M 0 when t=0. X 1M


*
0 is the theoretical initial

concentration of the reactant which will be entirely converted to 2-methylnaphthalene. It should

73
be equal to Xf, the ultimate yield of 2-methylnaphthalene, before decomposition occurs. The

analytical result of product yield is obtained with reference to Fogler’s book 27,

k1 X 1*M 0 −k1t
X 2M = [ e − e − k 2t ] (10)
k 2 − k1
Ea1 Ea2
where k1 = A1e RT and k 2 = A2 e RT .

Since the reaction temperatures and times are known, and A2 and Ea2 can be obtained from the

experimental study of 2-methylnaphthalene pyrolysis, there are three parameters to be estimated:

*
X 1M 0 , A1 and Ea1. With sufficient experimental data, the three kinetic parameters of 2-

methylnaphthalene formation in 1-methylnaphthalene pyrolysis can be estimated through the

nonlinear fitting method. The MATLAB source codes are shown in Appendix B. The results are

compared with the experimental data and are presented in Figure 5-4, which shows that the

model fits very well with the experimental data of the isomerization process.

74
16
2-methylnaphthalene formation in
1-methylnaphthalene pyrolysis

Yield (% carbon input)


12

0
800 850 900 950 1000

12
1-methylnaphthalene formation in
2-methylnaphthalene pyrolysis
Yield (% carbon input)

0
800 850 900 950 1000
Temperature (deg C)

Figure 5-4 Comparison between experimental data and modeling results for the isomerization
process in the pyrolysis of methylnaphthalenes. “Δ” denotes experimental data of 2-
methylnaphthalene formation from 1-methylnaphthalene pyrolysis, and “×” denotes those of 1-
methylnaphthalene formation from 2-methylnaphthalene pyrolysis. Solid lines denote modeling
results. Dash lines mark 95% confidence interval half-widths of prediction.

The calculated kinetic parameters for the series of reactions are shown in Table 5-6. The

activation energies for isomerization are much higher than are those for reactant decomposition

which are listed in Table 5-4. This is consistent with the assumption that only a fraction of the

reactant underwent isomerization.

Table 5-6 Kinetic Parameters obtained from the model for isomerization in pyrolysis of
methylnaphthalenes

75
Ea

Reaction lnA (kJ/mol) Xf (%)

1-methylnaphthalene formation in 2-

methylnaphthalene pyrolysis 34.20±16.47 337.40±162.36 16.86±4.37

2-methylnaphthalene formation in 1-

methylnaphthalene pyrolysis 35.79±4.37 362.48±42.89 31.38±2.11

5.5 Conclusion

This chapter investigated the kinetic properties of methylnaphthalene pyrolysis. The laminar

flow pattern in the reactor has been characterized as plug flow and it is proved that the error of

kinetic parameters Ea and lnA caused by the assumption is less than 12%. Global kinetic analysis

has been conducted on the consumption of 1- and 2-methylnaphthalene and their isomerization

processes with a series reaction model. Ea and lnA are 224.09±32.70 kJ/mol and 22.9±3.4 for 1-

methylnaphthalene, 262.57±76.26 kJ/mol and 26.6±7.7 for 2-methylnaphthalene. The simulated

temperature series of reactant consumption and product formation (isomerization) is consistent

with our experimental results. The kinetic parameters of 1-methylnapthalene pyrolysis are close

to the obtained values in other’s study on 1-methylnaphthalene pyrolysis in a batch reactor.

Hydrogen concentration during 1-methylnaphthalene pyrolysis was estimated and very low.

76
Reference
1. Skjoth-Rasmussen, M.; Glarborg, P.; Ostberg, M.; Johannessen, J.; Livbjerg, H.; Jensen,
A.; Christensen, T., Formation of polycyclic aromatic hydrocarbons and soot in fuel-rich
oxidation of methane in a laminar flow reactor. Combust Flame 2004, 2004, (136), 91-128.

2. Dagaut, P.; Ristori, A.; Bakali, A. E.; Cathonnet, M., Experimental and kinetic modeling
study of the oxidation of n-propylbenzene. Fuel 2002, 81, 173-184.

3. Burnham, A. K.; Braun, R. L., Global Kinetic Analysis of Complex materials. energy &
fuels 1999, 13, (1), 1-22.

4. Ledesma, E. B.; Marsh, N. D.; Sandrowitz, A. K.; Wornat, M. J., Global Kinetic Rate
Parameters for the Formation of Polycyclic Aromatic Hydrocarbons from the Pyrolyis of
Catechol, A Model Compound Representative of Solid Fuel Moieties. Energy & Fuels 2002, 16,
1331-1336.

5. Yang, J.; Lu, M., Thermal Growth and Decomposition of Methylnaphthalenes. Environ.
Sci. Technol. 2005, 39, 3077-3082.

6. Ramayya, S. V.; M.J Antal, J., Evaluation of systematic error incurred in the plug flow
idealization of tubular flow reactor data. energy & fuels 1989, 3, (1989), 105-108.

7. Cutler, A. H.; Antal, M. J.; Jones, M., A Critical Evaluation of the Plug-flow Idealization
of Tubular-flow Reactor Data. Ind. Eng. Chem. Res. 1988, 27, 691-697.

8. Kershenbaum, L. S.; Leaney, P. W., PYROLYSIS OF HYDROCARBONS IN A


LARGE PILOT-SCALE REACTOR. 2. PYROLYSIS OF n-BUTANE. Ind. Eng. Chem. Process
Des. Dev. 1986, 25, (3), 786-794.

9. Behar, F.; Budzinski, H.; Vandenbroucke, M.; Tang, Y., Methane Generation from Oil
Cracking: Kinetics of 9-Methylphenanthrene Cracking and Comparison with Other Pure
Compounds and Oil Fractions. Energy & fuels 1999, 13, 471-481.

10. Smith, C. M.; Savage, P. E., Reactions Of Polycyclic Alkylaromatics .4. Hydrogenolysis
Mechanisms In 1-Alkylpyrene Pyrolysis. Energy & Fuels 1992, 6, (2), 195-202.

11. Leininger, J.-P.; Lorant, F.; Minot, C.; Behar, F., Mechanisms of 1-Methylnaphthalene
Pyrolysis in a Batch Reactor. energy & fuels 2006, 20, (6), 2518-2530.

12. Furue, H.; Pacey, P. D., Performance of Cylindrical Flow Reactors in a Kinetic Study of
the Isomerization of Cyclopropane. J. Phys. Chem. 1980, 84, 3139-3143.

77
13. Cussler, E. L., Diffusion Mass Transfer in Fluid Systems (2nd Edition). Cambridge
University Press: Cambridge, United Kingdom, 1997.

14. Kimura, Y.; Abe, D.; Terazima, M., Vibrational Energy Relaxation of Naphthalene in the
S1 State in Various Gases. J. Chem. Phys. 2004, 121, (12), 5794-5800.

15. Yaws, C. L., Handbook of Viscosity V. 4. Inorganic Compounds and Elements. Gulf Pub.
Co.: Houston, Texus, 1995; Vol. 4.

16. Walker, R. E., Chemical reaction and diffusion in a catalytic tubular reactor. The physics
of fluids 1961, 4, (10), 1211-1216.

17. Dang, V. D.; Steinberg, M., Convective Diffusion with Homogeneous and Heterogeneous
Reactions in a Tube. J. Phys. Chem. 1980, 84, 214-219.

18. Cleland, F. A.; Wilhelm, R. H., Diffusion and Reaction in viscous-flow tubular reactor.
AIChE Journal 1956, 2, 489-497.

19. Saxena, S. C., Devolatilization and Combustion Charateristics of Coal Particles. Prog.
Energy Combust. Sci. 1990, 16, 55-94.

20. Yang, J.; Lu, M., One Improved Global Kinetic Analysis Based on Combustion
Experiments at Different Temperatures. In 4th Joint Meeting of the U. S. Sections of the
Combustion Institute, Philadelphia, PA, 2005.

21. Berthouex, P. M.; Brown, L. C., Statistics for Environmental Engineers (2nd Edition).
Lewis Publishers: Boca Raton, Florida, 2002; p 397.

22. Lapin, L. L., Probability and Statistics for Modern Engineering (2nd Edition). PWS-
KENT: Boston, 1990.

23. Cvetanovic, R. J.; Singlelon, D. L.; Paraskevopoulos, G., Evaluations of the Mean Values
and Standard Errors of Rate Constants and Their Temperature Coefficients. J. Phys. Chem. 1979,
83, (1), 50-60.

24. Dang, V. D., Steady-State Mass Transfer with Homogeneous and Heterogeneous
Reactions. AIChE Journal 1983, 29, (1), 19-25.

25. Shaddix, C. R.; Brezinsky, K.; Glassman, I., Analysis of Fuel Decay Routes in the High-
Temperature Oxidation of 1-Methylnaphthalene. Combust. Flame 1997, 108, 139-157.

26. Pitsch, H., Detailed Kinetic Reaction Mechanism for Ignition and Oxidation of alfa-
Methylnaphthalene. In 26th Symp. Int. Combust., The Combustion Institute: 1996; pp 721-728.

78
27. Fogler, H. S., Elements of Chemical Reaction Engineering. Prentice-Hall: Englewood
Cliffs, New Jersey, 1986.

79
6 CHAPTER VI ETHYLNAPHTHALENE PYROLYSIS

6.1 Introduction

Compared with the number of studies on naphthalene 16, 30, 67, 86, 87 and methylnaphthalene 57, 58, 70,

87, 88
pyrolysis, the study on the pyrolysis of 1- and 2- ethylnaphthalene has been insufficient.

Studies on methylnaphthalenes have indicated that in addition to the formation of

naphthylmethyl radical, hydrogen displacement of the methyl site chain can also contribute to

PAH and soot formation70. The methyl displacement of the site chain can even be favored at

higher temperatures over that of benzylic radical formation. In order to better understand the

reactivity of the site chain, it is necessary to study ethylnaphthalenes, which have a longer side

chain than methylnaphthalenes and can lead to more complex product formation mechanisms.

Beltrame et al.89, 90 studied the hydrodealkylation of 1 and 2-ethylnaphthalene from 550-710°C at

elevated pressures (2.9 to 44.6 atm). The main products of 2-ethylnaphthalene hydrodealkylation

were naphthalene, 2-methylnaphthalene and 2-vinylnaphthalene, whose total yields accounted

for more than 96% of naphthalene ring input (the author’s approach to account for product

yields)89. The rapidly decreasing yield of 2-vinylnaphthalene at pressures higher than 20 atm

served as an indication that it is the product of dehydrogenation. The identified gaseous products

were methane, ethane, and ethylene, in the order of abundance. In the subsequent study of 1-

ethylnaphthalene hydrodealkylation, naphthalene, 1-methylnaphthalene, acenaphthene and

acenaphthalene were identified as main products accounting for more than 97% of naphthalene
80
ring input. The pathways of 1-ethylnaphthalene forming 1-methylnaphthalene, naphthalene and

acenaphthalene were proposed in the kinetic study, and kinetic parameters were obtained from

experimental data.90 The reactivity of the ethyl side chain has been observed in this study,

however, the high concentrations of hydrogen and high pressure are not always typical in most

combustion applications. The study of 1-ethylnaphthalene pyrolysis at atmospheric pressure is

still essential in understanding its roles in practical combustion conditions.

The experimental study of 1-ethylnaphthalene pyrolysis is presented in this chapter. Only one

ethylnaphthalene isomer was studied as our earlier work on 1 and 2-methylnaphthalenes

indicated that the governing thermal decomposition mechanisms of the two methylnaphthalenes

are similar. Pyrolytic products have been identified and their possible formation mechanisms are

discussed. The global kinetic analysis for 1-ethylnaphthalene consumption is performed based on

experimental results.

6.2 Experimental Setup

1-Ethylnaphthalene (98% with 1.94% 2-ethylnaphthalene and 0.06% 1-(2-propyenyl)-

naphthalene) was used and was purchased from Aldrich. The input reactant concentrations and

gas flow rates are listed in Table 6-1.

81
Table 6-1 Initial experimental conditions

Temperature (ºC)
Parameters
600 650 700 750 800 850
Helium Flow rate (nlpma) 2.8 2.65 2.52 2.4 2.28 2.18
Reactant input
416 439 462 485 510 534
Concentration (ppm)

a
nlpm: normal liters per minute at 20 ºC, 1 atm.

6.3 Results and Discussion

6.3.1 Experiment results

Figure 6-1 gives an example of the total ion chromatogram of identified products from 1-

ethylnaphthalene pyrolysis at 850ºC, with all the major products labeled. Some visible unlabelled

peaks are from the septum bleed. The experimental data of product yields are shown in Table B-

4.

Figure 6-1 Total Ion chromatogram of 1-ethylnaphthalene pyrolysis at 850 ºC

82
Figure 6-2 presents the temperature series of 1-ethylnaphthalene pyrolysis. Average values are

used for each data point, and the error bars represent the actual maximum and minimum values.

The carbon recovery rate is nearly 100% throughout the temperature range, which suggests that

the products have been effectively collected and identified. At 600ºC, more than 90% of 1-

ethylnaphthalene remains unreacted, while almost all is decomposed at 850ºC. This indicates that

the selected temperature range is appropriate for studying 1-ethylnaphthalene pyrolysis. Soot is

not observed except in negligible amounts at 850ºC. This may be due to the lower temperature

range for 1-ethylnaphthalene pyrolysis compared with that for methylnaphthalenes (800 to

1000ºC )70, which is an indication that 1-ethylnaphthalene may be more reactive than the

methylnaphthalenes.

120

100
Yield (% carbon input)

80

60 1-EN
Recovery
PAH
40

20

0
600 650 700 750 800 850
Temperature (deg C)

Figure 6-2 Overall temperature series of 1-ethylnaphthalene pyrolysis. “1-EN”: the quantity of
unconsumed 1-ethylnaphthalene; “PAH”: the total yields of all of aromatic products identified;
“Recovery”: the total recovery, which is the sum of “1-EN” and “PAH”.

83
Figure 6-3 presents the major product yields. The identified products with yields greater than 1%

include 1-vinylnaphthalene, 1-methylnaphthalene, acenaphthalene, naphthalene, 2-

methylnaphthalene, 2-vinylnaphthalene, and acenaphthene. The most abundant product is 1-

vinylnaphthalene and its yield decreases slightly when the temperature is higher than 750ºC. The

second most abundant is acenaphthene at temperatures below 700ºC and is 1-methylnaphthalene

at higher temperatures. The minor products (with less than 1% yield) identified included indene,

1,7-, 1,6-, 1,4- and 1,5-dimethylnaphthalene, fluorene, phenanthrene, and benzofulvene. Most of

these minor products were only found at temperatures higher than 750ºC.

100

10
Yield (% carbon input)

1-VN
0.1 1-MN
AceN
naphthalene
0.01 2-MN
2-VN
acenaphthene
0.001
600 650 700 750 800 850
Temperature (deg C)

Figure 6-3 Major products and their yields in 1-ethylnaphthalene pyrolysis. “VN” represents
vinylnaphthalene; “MN” represents methylnaphthalene; “AceN” represents acenaphthalene.

Minor quantities of gaseous products have also been quantified. The gas sample at 800ºC was

analyzed using GC-FID by our colleague at Ohio University. The instrument was equipped with

a cryogenic pre-concentration system. Ethane and ethylene were identified and each accounted

for approximately 1.8% carbon input. However, the system was not setup for methane analysis.
84
Methane concentrations at 700 and 850ºC were measured by a multiple gas analyzer based on

Fourier transform infrared (FTIR) Spectroscopy and the yields were 0.4% and 3.6% respectively.

Most gas molecules can absorb infrared radiation with a characteristic, molecule-specific

absorption spectrum except noble gases and homonuclear diatoms which are infrared inactive.

Gas species can be identified and quantified through the location and magnitude of these

absorptions.

6.3.2 Postulated reaction mechanisms

CH3 CH2
(1)
+ H

CH3 CH3
HC
(2)
+ H

CH3
CH2

(3)
+ CH3

CH3

+ CH3CH2 (4)

85
CH3 CH2
(5)
+ H + H2

CH3 CH3
HC
(6)
+ H + H2

CH3
CH2

+ H
+ CH4 (7)

CH3

+ H + C2H 5
(8)

The initiation of 1-ethylnaphthalene pyrolysis can occur via four possible pathways as described

by reactions (1) to (4). Reactions (1) and (2) involve the dissociation of the C-H bond at the

primary and benzylic locations. Reactions (3) and (4) are resultant from the C-C bond scission at

the benzylic site and the aromatic ring respectively. One study on ethylbenzene pyrolysis44

suggested that the C-H homolysis is similar to reaction (2) as the major reaction mechanism,

while reaction (3) was regarded as the major pathway by other studies42, 43, 45, 46.

The heats of reaction of the four pathways were estimated and listed in Table 6-2, based on

thermodynamic data obtained from the literature91 and the NIST Chemistry WebBook

(http://webbook.nist.gov/chemistry/), as an indication of the relative importance of these steps in

product formation. Our results indicated that both reactions (2) and (3) are favored energetically

due to the formation of resonance stabilized radicals, with (3) being slightly more important.

These results qualitatively agree with McMillen et al.92. Reaction (4) is the least favored as it
86
disturbs the aromatic structure. As the reactions proceed, the radical pools build up, and those

initiation steps give way to hydrogen abstractions expressed by reactions (5) to (8) and radical

displacement pathways due to their high activation energies. It should be noticed that in reaction

(8) naphthalene and ethyl radical are formed instead of naphthyl radical and ethane since the

former has a lower heat of reaction -45.38 kJ mol-1 than -2.93 kJ mol-1, that of the latter pathway.

And the formation of naphthylethyl radicals is still favored thermodynamically.

Table 6-2 Heats of reaction (kJ mol-1) for the reactant consumption pathways in 1-EN pyrolysis

Reaction 1 2 3 4 5 6 7 8

Calculated 413.62 341.31 322.56 417.87 -22.376 -94.69 -117.31 -45.38


92
355.7 304.7

Figure 6-4 describes the hypothesized reaction mechanisms of 1-ethylnaphthalene pyrolysis.

Hydrogen abstraction from the ethyl group results in naphthylethyl radicals at benzylic and

primary locations. Both of these steps result in the formation of 1-vinylnaphthalene, with the

hydrogen abstraction at the benzylic site being favored. Hydrogen displacement of the ethyl

group gives rise to naphthalene, and the homolytic cleavage of the ethyl group formed 1-

methylnaphthalene. These results are in agreement with the pathways postulated for

ethylbenzene oxidation93.

87
-2H

acenaphthene acenaphthalene

-2H
-H +C
2H 2 -2H

CH3 CH2 CH2


HC
CH2
-H (isomerize)
and
2-vinylnaphthalene
1-vinylnaphthalene
+ H -H2 -C2H4 +H -C2H4 + 2H
or -H

CH3 CH2

+H -C2H5 (isomerize) -CH2 + 2H

indene
naphthalene benzofulvene

-CH3
-H+CH3
+ H -CH3

CH2 CH3
CH3
+H (isomerize)

1-methylnaphthalene 2-methylnaphthalene

Figure 6-4 Proposed reaction mechanisms of 1-ethylnaphthalene pyrolysis.

The yields of 1-vinylnaphthalene increase with temperature until 750ºC, and its decrease can be

the result of isomerization to form 2-vinylnaphthalene or formation of naphthalene by hydrogen

displacement, which is analogous to that of styrene93. In addition, the cyclodehydrogenation of 1-

vinylnaphthalene gives rise to acenaphthalene. Similarly, acenaphthene can be resultant from the

cyclodehydrogenation of the naphthylethyl radical and further dehydrogenation produces

acenaphthalene. The hypothesized pathway of acenaphthene and acenaphthalene formation is in

agreement with the study on hydrodealkylation of ethylnaphthalenes90, and is also supported by

the experimental results in Figure 6-3. Also, the addition of C2H2 to naphthalene or naphthyl

radical can be another possible route of acenaphthalene formation, as described in 1-

88
methylnaphthalene pyrolysis70. Due to the higher activation energy involved in the C-C

homolysis of the ethyl group, it tends to be favored at higher temperatures, which is evidenced

by the temperature series of 1-methylnaphthalene in Figure 6-3.

Naphthalene can be produced via multiple pathways, such as hydrogen displacement of the side

chain from 1-ethylnaphthalene and 1-vinylnaphthalene or as a product of 1-methylnaphthalene

pyrolysis as observed in our 1-methylnaphthalene study. However, hydrogen displacement of the

methyl group on 1-methylnaphthalene should not be the major route for naphthalene formation

in this study.

The maximum possible yield of naphthalene from the 1-methylnaphthalene pathway is estimated

as the multiplication product of 1-methylnaphthalene formed in 1-ethylnaphthalene pyrolysis

and the conversion factor of 1-methylnaphthalene to naphthalene obtained from an earlier study70.

The maximum yield of 1-methylnaphthalene is shown in Figure 6-5, with the assumption that 1-

ethylnaphthalene pyrolysis (the consumed 1-ethylnaphthalene) only resulted in 1-

vinylnaphthalene, acenaphthene and 1-methylnaphthalene, while the other products are

secondary products from these three compounds.

70
In 1-methylnaphthalene pyrolysis , the yield of naphthalene is less than 2.4% at 800ºC and

reaches 5.3% at 850ºC. Therefore, the maximum possible yield of naphthalene from the 1-

89
methylnaphthalene pathway is 0.87% at 800ºC and 2.7% at 850ºC, which are much lower than

the actual experimental values (4.0% at 800ºC and rises to 8.7% at 850ºC).

120
residual 1-EN
consumed 1-EN
100 1-VN + AceN
Yield (% carbon input)

maximum 1-MN

80

60

40

20

0
600 650 700 750 800 850
Temperature (deg C)

Figure 6-5 The estimated maximum yield of 1-methylnaphthalene. “residual 1-EN”: yields of
unconsumed 1-ethylnaphthalene; “consumed 1-EN”: quantity of consumed 1-ethylnaphthalene,
which is 1-“residual 1-EN”; “1-VN + AceN”: the sum of 1-vinylnaphthalene and acenaphthalene;
“maximum 1-MN”: the estimated maximum possible yield of 1-methylnaphthalene in 1-
ethylnaphthalene pyrolysis.

The identification of 2-methylnaphthalene and 2-vinylnaphthalene in 1-ethylnaphthalene

pyrolysis can be an indication of the isomerization process70. In addition, the higher yield of 2-

methylnaphthalene from 1-ethylnaphthalene pyrolysis (1.1% at 800ºC and 5.5% at 850ºC) than

that from 1-methylnaphthalene pyrolysis (0.8% at 800ºC and 2.3% at 850ºC) indicates that

isomerization might not be the only route of 2-methylnaphthalene formation. Another possible

pathway may be from methyl addition to naphthalene, which agrees with the study of Mimura et

al.86.
90
Less than 2% 2-ethylnaphthalene, which should be impurity in the reactant, was identified in

products at different temperatures and its yield gradually decreased from 2% at 600°C to 0.1% at

850°C. Therefore, isomerization didn’t exist or at least was insignificant in this study. One

postulation is that the isomerization process, the H addition and intramolecular alkyl shift, as

described in our 1-methylnaphthalene pyrolysis, has a much higher energy barrier than other

reactions. Our future work on 1-propylnaphthalene is underway to further investigate the

isomerization process.

6.3.3 Global kinetic study on 1-ethylnaphthalene consumption

The global kinetic study can provide an overall estimate of the reactant consumption rate in 1-

ethylnaphthalene pyrolysis.19 The kinetic data, such as the pre-exponential factor and the

activation energy, can be obtained through the Arrhenius equation, and can be used to evaluate

the reactivity of different reactants. In our experimental system, the concentration of residual

reactant after time t at the temperature T can be expressed as 94

Ea
X = X 0 exp(− At exp(− )) (3)
RT

where A is the pre-exponential factor (s-1), and Ea is the activation energy (kJ mol-1). R is the

universal gas constant. With the nonlinear least squares regression function “nlinfit” in Matlab

7.1, global kinetic parameters A = e15.2 and Ea=125.9 kJ mol-1 are obtained with the given

91
experimental data of yields and temperatures.. The calculated curve and original experimental

data are displayed in Figure 6-6 which shows that the model result is in good agreement with

experimental data. Similarly, the activation energies for 1-methylnaphthalene and 2-

methylnaphthalene have been calculated using experimental data, which are 224.1 kJ mol-1 and

262.6 kJ mol-1 respectively94. This suggests the reactivity order of three compounds: 1-

ethylnaphthalene>1-methylnaphthalene>2-methylnaphthalene, which is consistent with our

experimental observations.

100
Yield (% carbon input)

80

60

40

20

0
600 700 800 900
Temperature (deg C)

Figure 6-6 The experimental (hollow diamond points) and theoretical results (curve) of 1-
ethylnaphthalene decomposition.

6.4 Conclusion

The products of 1-ethylnaphthalene pyrolysis suggest the following reaction pathways: hydrogen

abstraction from the ethyl group results in naphthylethyl radicals at benzylic and primary

locations, both lead to the formation of 1-vinylnaphthalene, with the benzylic route being

favored; hydrogen displacement of the ethyl group results in naphthalene; and the homolytic
92
cleavage of the ethyl group gives rise to 1-methylnaphthalene. The relative abundance of these

products is also an indication of the relative importance of these pathways. The reactivity of 1-

ethylnaphthalene is higher than that of methylnaphthalenes, as evidenced by the lower reaction

temperatures, and the higher reactant consumption/product formation at the same temperature.

The experimental results of 1-ethylnaphthalene pyrolysis fits very well with a first order reaction

model, and the activation energy Ea and Arrhenius preexponential factor A were determined as

125.9 kJ mol-1 and e15.2 respectively. Results of the global kinetic study are also indicative of the

higher reactivity of 1-ethylnaphthalene than 1-methylnaphthalene, as evidenced by the much

lower activation energy (224.1 kJ mol-1 for 1-methylnaphthalene).

References
1. Tesner, P. A.; Shurupov, S. V., Soot Formation during Pyrolysis of Naphthalene,
Anthracene and Pyrene. Combust. Sci. Technol 1997, 126, 139-151.

2. Lu, M.; Mulholland, J. A., PAH Growth from the Pyrolysis of CPD, Indene and
Naphthalene Mixture. Chemosphere 2004, 55, 605-610.

3. Badger, G. M.; Jolad, S. D.; Spotswood, T. M., The Formation of Aromatic


Hydrocarbons at High Temperatures: xx. The Pyrolysis of [1-14C]Naphthalene. Aust. j. chem.
1964, 17, 771-777.

4. Mimura, K.; Madono, T.; Toyama, S.; Sugitani, K.; Sugisaki, R.; Iwamatsu, S.; Murata,
S., Shock-induced Pyrolysis of Naphthalene and Related Polycyclic Aromatic Hydrocarbons
(Anthracene, Pyrene, and Fluoranthene) at Pressures of 12-33.7 GPa. J. Anal. Appl. Pyrolysis
2004, 72, 273-278.

5. Platonov, V. V.; Ivleva, L. N.; Vol-Epshtein, A. B.; Kasimtseva, T. V.; Klyavina, O. A.,
Kinetics of the Homogenous Pyrolysis of Naphthalene and Monomethylnaphthalenes. Solid Fuel
Chemistry (English Translation of Khimiya Tverdogo Topliva) 1986, 20, 74-82.

93
6. Lijinsky, W.; Raha, C. R., The Pyrolysis of 2-Methylnaphthalene. J. org. chem 1961, 26,
3566-3568.

7. Shaddix, C. R.; Brezinsky, K.; Glassman, I., Analysis of Fuel Decay Routes in the High-
Temperature Oxidation of 1-Methylnaphthalene. Combust. Flame 1997, 108, 139-157.

8. Shaddix, C. R.; Brezinsky, K.; Glassman, I., Oxidation of 1-Methylnaphthalene. In 24th


Symp. Int. Combust., The Combustion Institute: 1992; pp 683-690.

9. Yang, J.; Lu, M., Thermal Growth and Decomposition of Methylnaphthalenes. Environ.
Sci. Technol. 2005, 39, 3077-3082.

10. Beltrame, P.; Carnltl, P.; Maronglu, B.; Mura, L.; Solinas, V.; Mori, S.,
Hydrodealkylation of 2-Ethylnaphthalene. A Study of Reaction Products and Kinetics. Ind. Eng.
Chem. Process Des. Dev. 1979, 18, 338-342.

11. Beltrame, P.; Carnltl, P.; Maronglu, B.; Mura, L.; Solinas, V.; Mori, S.,
Hydrodealkylation of 1-Ethylnaphthalene and its Side Reactions. A Kinetic Study. Ind. Eng.
Chem. Process Des. Dev. 1981, 20, (379-384).

12. Brouwer, L.; Muller-Markgraf, W.; Troe, J., Identification of Primary Reaction Products
in the Thermal Decomposition of Aromatic Hydrocarbons. In 20th Symp. Int. Combust., The
Combustion Institute: 1984; p 799.

13. Davis, H. G., Rate of formation of toluene from ethylbenzene. Int. J. Chem. Kinet. 1983,
15, 469-474.

14. Robaugh, D. A.; Stein, S. E., Very-low-pressure Pyrolysis of Ethylbenzene,


Isopropylbenzene, and Tert-butylbenzene. Int. J. Chem. Kinet. 1981, 13, 445 -462.

15. Pamidimukkala, K. M.; Kern, R. D., The High Temperature Pyrolysis of Ethylbenzene.
Int. J. Chem. Kinet. 1986, 18, 1341-1353.

16. Mizerka, L. J.; Kiefer, J. H., The High-temperature Pyrolysis of Ethylbenzene Evidence
for Dissociation to Benzyl and Methyl Radicals. Int. J. Chem. Kinet. 1986, 18, 363-378.

17. Curran, H.; C.Wu; Marinov, N.; Pitz, W. J.; Westbrook, C. K.; Burcat, A., The Ideal Gas
Thermo-dynamics of Diesel Fuel Ingredients. I. Naphthalene Derivatives and their Radicals. J.
Phys. Chem. Ref. Data 2000, 29, 463-517.

18. McMillen, D. F.; Golden, D. M., Hydrocarbon Bond Dissociation Energies Ann. Rev.
Phys. Chem. 1982, 33, 493-532.

94
19. Litzinger, T. A.; Brezinsky, K.; Glassman, I., The Oxidation of Ethylbenzene near 1060
K. Combust Flame 1986, 63, 251-267.

20. Burnham, A. K.; Braun, R. L., Global Kinetic Analysis of Complex materials. energy &
fuels 1999, 13, (1), 1-22.

21. Yang, J.; Lu, M., Global Kinetic Analysis of Pyrolysis of Methylnaphthalenes. Appl.
Anal. Pyrolysis 2006.

95
7 CHAPTER VII 1-PROPYLNAPHTHALENE PYROLYSIS

7.1 Introduction

In our previous experiments, pyrolysis of 1- and 2-methylnaphthalene and 1-ethylnaphthalene

was studied.70, 94, 95


Isomerization was observed in pyrolysis of methylnaphthalenes as a

significant product formation pathway and absent in 1-ethylnaphthalene pyrolysis, which showed

that the length of the side chain of alkylated naphthalene is an important factor in pyrolysis. It is

possible that the longer the alkyl chain, the less isomerization happens. For better understanding

of the effect of the alkyl chain on PAH formation during combustion, pyrolysis of 1-

propylnaphthalene, an alkylated naphthalene with a longer chain, is investigated experimentally

and kinetically. The same flow reactor system and experimental method that were used in

previous studies are also employed in this study for consistency. In addition, more experiments

with 1.9 second retention time were conducted to study the influence of retention time, which

was not investigated in our former studies. Pyrolysis products were measured quantitatively and

their yields as a function of reaction temperature are shown. The routes of reactant

decomposition and further radical reactions are discussed based on product information.

Furthermore, the hypothesized reaction mechanism based on those routes is demonstrated. Using

the global kinetic model of reactant consumption and product formation introduced in the former

study94, 1-propylnaphthalene decomposition and formation of 1- and 2-methylnaphthalene and 1-

96
ethylnaphthalene are studied kinetically. The effects of temperature and retention time on yields

are discussed.

7.2 Experimental Design

The experimental setup for 1-propylnaphthalene pyrolysis is the same as that used in 1- and 2-

methylnaphthalene combustion and 1-ethylnaphthalene pyrolysis. The experimental and

analytical procedures were introduced in detail elsewhere70 and is briefly described below.

The 1-propylnaphthalene (99.2% with 1-ethylnaphthalene, 1-hexylnaphthalene and 1-

propenylnaphthalene as impurity), whose density at 20 °C is 0.9897 kg/liter96, was synthesized

by Chiron AS, Norway. In each experiment, 30μl reactant was injected evenly by a syringe pump

into a glass vaporizer heated to approximately 200ºC. The reactant was vaporized quickly and

the vapor was delivered into the 17mm ID quartz tube reactor at isothermal conditions by the

carrier gas, helium. The experiment temperature ranged from 550 to 750ºC with 50ºC increments

and the retention time was fixed to 1 and 1.9 seconds to study the effects of both temperature and

reaction time on the pyrolysis. The initial reactant concentrations and gas flow rates for one

second are listed in Table 7-1. Each run lasted for four minutes, plus two more minutes to flush

the residual reactant and products in the reactor. Multiple runs were performed for each

experimental condition. Products were collected by an ice bath solvent trap filled with

dichloromethane, and quantified by a Varian Saturn 2200 GC/MS with a CP-Sil 8 CB Low

Bleed/MS capillary column (30m × 0.25mm × 0.25μm) using modified EPA method 525. Detailed
97
parameters of instrument setup have been described in the preliminary study95, 97. Product yields

are reported as percent of input on carbon basis in this chapter.

Table 7-1 Initial experimental conditions for 1 second

Temperature (ºC)
Parameters
550 600 650 700 750

Helium Flow rate (nlpma) 2.98 2.8 2.65 2.52 2.4

Reactant input
352 374 396 416 437
concentration (ppm)

a
nlpm: normal liters per minute at 20ºC, 1 atm.

7.3 Results and Discussion

7.3.1 Experimental results

One chromatogram of products from 1-propylnaphthalene pyrolysis at 700ºC is shown in Figure

7-1. In total, there are 26 species most of which are identified by comparing their spectra and

retention times with those of standard solutions or a commercially provided spectrum library. If

the chemical structure could not be decided, the molecular formula was used instead. Five

compounds eluting 20 minutes after the sample being injected are shown in the small zone.

Those unlabelled peaks are signals of septum bleed which are silicone containing compounds.

The experimental data of product yields are shown in Table B-5 for 1 second and Table B-6 for

98
1.9 second. The chemical structures of some products are shown and discussed in subsequent

sections.

1.00 6 8
17
MCounts

18

9 10
14

0.75

0.50

7
13
0.25 16
12

19
21
11
5
2 3
1 4 15 20
0.00
5.0 7.5 10.0 12.5 15.0 17.5
minutes

Figure 7-1 A chromatogram of 1-propylnaphthalene pyrolysis at 700°C. 1.toluene, 2. phenylethyne,


3. indene, 4. C10H10, 5. benzofulvene, 6. naphthalene, 7. 2-methylnaphthalene, 8. 1-
methylnaphthalene, 9. 1-ethylnaphthalene, 10. 1-vinylnaphthalene, 11. 1,7-dimethylnaphthalene, 12.
2-vinylnaphthalene, 13. 14-dimethylnaphthalene, 14. acenaphthalene, 15. 12-dimethylnaphthalene,
16. 1-allylnaphthalene, 17. 1-propylnaphthalene, 18. acenaphthene, 19. 1-isoallylnaphthalene, 20.
fluorene, 21. C13H12, 22. phenanthrene, 23. 1-phenylmethyl-naphthalene, 24. 1,1-binaphthyl, 25. 1,
2’-dinaphthylmethane, 26. 1, 2-di-1-naphthylethane

In Figure 7-2, the yield of residual 1-propylnaphthalene and aromatic products, which are

defined as all of the observed aromatic compounds in products except the reactant, is shown for

both the one and 1.9 second retention times. Average values are used and the error bar of 1-

propylnaphthalene represents the actual maximum and minimum values that were measured. The

residual reactant decreases from 97.8% at 550ºC to 0.03% at 750ºC for one second and from
99
80.9% at 500ºC to almost zero at 750ºC for 1.9 seconds. Therefore, the effect of temperature on

the decomposition of 1-propylnaphthalene can be clearly demonstrated, and the reactions

happening during pyrolysis can be studied in detail in the selected temperature range. Figure 7-2

also shows that more reactant was consumed at the same temperature, and more products were

formed correspondingly when the reaction time changed from one second to 1.9 seconds.

According to the study of PAH formation in catechol pyrolysis by Marsh et al. 98, the effect of

reaction or residence time on the yields of specific products changed with different pyrolysis

temperatures. In that study, the yields of 7 most prominent PAHs increased with increasing

retention time from 0.4 to 1 second at 800°C. While at 1000°C, their yields all decreased. The

detailed kinetic analysis and further discussion of the reactant decomposition and formation of

some products will be performed later to explain the residence time and temperature effect. The

carbon recovery of two sets of experiments differs from 67% to 101%. The low carbon recovery

may be due to the formation of gaseous products. Since 1-propylnaphthalene has a longer side

chain, the decomposition or removal of its side chain could give rise to the formation of more

gaseous products than that of methylnaphthalenes and ethylnaphthalene. And gaseous products

are not counted when calculating carbon recovery, which may be the major reason of low carbon

recoveries in this study.

100
100

Yield (% carbon input)


80

60 1PN (1 sec)
Aromatic (1 sec)
1PN (1.9 sec)
40 Aromatic (1.9 sec)
Carbon recovery (1 sec)
Carbon recovery (1.9 sec)
20

0
500 550 600 650 700 750 800
Temperature (deg C)

Figure 7-2 The yield of residual 1-propylnaphthalene and total aromatic products as a function of
time and temperature. “1PN” represents 1-propylnaphthalene. “carbon recovery” represents the
sum of residual reactant, aromatic products and soot on carbon basis.

The major products whose yields are larger than 1% include 1-vinylnaphthalene, 1, 2-di-1-

naphthylethane, 1-methylnaphthalene, 1-ethylnaphthalene, naphthalene, and acenaphthalene.

Their yields as the function of temperature for the 1 second retention time are shown in Figure 7-

3. When the temperature reaches 750ºC, the yield of 1, 2’-dinaphthylmethane is also higher than

1%. The most abundant product is 1-vinylnaphthalene. Its yield increases from 1.93% at 550ºC

to 36.6% at 700ºC and then decreases slightly to 34.7% at 750ºC. The second most abundant is 1,

2-di-1-naphthylethane up to 650ºC and 1-methylnaphthalene at higher temperatures. The yield of

1, 2-di-1-naphthylethane rises to its maximum value, 15.2% at 650ºC from 0.17% at 550ºC and

then lessens at higher temperatures.

101
100

Y ield (% ca rb on in pu t)
10

0.1 N aphthalene
1M N
1EN
1V N
0.01 12D N E
A ceN
12D N M

0.001
550 600 650 700 750 800
Te m p eratu re (de g C )

Figure 7-3 Major product yields of 1-propylnaphthalene pyrolysis for 1 second. “1MN” represents
1-methylnaphthalene; “1EN” represents 1-ethylnaphthalene; “1VN” represents 1-vinylnaphthalene;
“12DNE” represents 1, 2-di-1-naphthylethane; “AceN” represents acenaphthalene; “12DNM”
represents 1, 2’-dinaphthylmethane.

Most of those major products presented were also in the product list of 1-ethylnaphthalene

pyrolysis except one C22H18 and one C21H16. Even though we did not have standards of those two

compounds, both the result of library searching and the existence of abundant 1- alkylated

naphthalenes strongly suggest they are 1, 2-di-1-naphthylethane and 1, 2’-dinaphthylmethane

respectively. The mass spectra and chemical structures of those two compounds shown in Figure

7-4 suggest that those two compunds result from the combination of two naphthylmethyl radicals

and one naphthylmethyl radical with a naphthyl radical. The occurrence of 1, 2’-

dinaphthylmethane at temperatures higher than 700ºC may indicate the existence of a significant

amount of naphthyl radical, which was supported by that the concentration of naphthalene

reaches 4.1% at 700ºC from 0.36% at 600ºC in our experiments. It also provides explanations for

the absence of the two compounds in the product list of pyrolysis of methylnaphthalenes and 1-

ethylnaphthalene. In pyrolysis of methylnaphthalenes the reactants almost didn’t decompose and


102
no significant amount of naphthylmethyl radical was formed when the pyrolysis temperature was

below 800ºC. At higher temperatures, a large quantity of naphthyl radical was formed and

binaphthyls became major products. In 1-ethylnaphthalene pyrolysis, the major reactant

consumption in the temperature range was dehydrogenation on the side chain causing the

formation of 1-vinylnaphthalene eventually. Even though 1-methylnaphthalene yield reached

12.2% at 750ºC and even higher at higher temperatures, which is an indicator of naphthylmethyl

radical existence, 1, 2-di-1-naphthylethane is unstable and its formation should not be

remarkable, shown as its dramatically decreased yield at 750ºC in this study.

103
Scan 2735 from E:\Experimental Data\PN\7-17-06\data\PN700_7_3.SMS
Scan 2656 from E:\Experimental Data\PN\7-17-06\data\PN700_7_3.SMS
Purity = 142 Fit = 165 RFit = 208 Average = 172 Ion Range: 40 - 450
Spect 1
BP 141 (12719=100%) PN700_7_3.SMS 32.853 min. Scan: 2735 Chan: 1 Ion: 14108 us RIC: 30546 BC
141
100%

75%

50% 1, 2-di-1-naphthylethane
115
25% 282

63 89 191 207 265 327 343 429


0%
Spect 2
BP 268 (711=100%) PN700_7_3.SMS 31.941 min. Scan: 2656 Chan: 1 Ion: 23057 us RIC: 4849 BC
268
100%

75%

50%
1, 2'-dinaphthylmethane
115
25%
208
63 192 415
0%
100 200 300 400
m/z

Target Spectrum from e:\... \pn\7-17-06\data\pn700_7_3.sms


Scan No: 2735, Time: 32.853 minutes
No averaging. Background corrected.
Comment: 32.853 min. Scan: 2735 Chan: 1 Ion: 14108 us RIC: 30546 BC
Pair Count: 133 MW: 0 Formula: None CAS No: None Acquired Range: 40 - 450
Match Spectrum from e:\... \pn\7-17-06\data\pn700_7_3.sms
Scan No: 2656, Time: 31.941 minutes
No averaging. Background corrected.
Comment: 31.941 min. Scan: 2656 Chan: 1 Ion: 23057 us RIC: 4849 BC
Pair Count: 122 MW: 0 Formula: None CAS No: None Acquired Range: 40 - 450

Figure 7-4 The mass spectra and chemical structures of 1, 2-di-1-naphthylethane and 1, 2’-
dinaphthylmethane.

Other minor species observed include toluene, phenylethyne, indene, C10H10, benzofulvene, 2-

methylnaphthalene, 17-, 14-, and 12-dimethylnaphthalene, 2-vinylnaphthalene, acenaphthene, 1-

allylnaphthalene, 1-isoallylnaphthalene, unrecognized C13H12, fluorene, phenanthrene, 1-

phenylmethyl-naphthalene, and 1,1-binaphthyl. There is no isomer of 1-propylnaphthalene

observed in products, and the yields of 2-methylnaphthalene and 2-vinylnaphthalene are smaller

than those in 1-ethylnaphthalene experiments.


104
7.3.2 Postulated reaction mechanism

The major products of 1-propylnaphthalene pyrolysis, except acenaphthalene, are very analogous

to those of the combustion of 1-propylbenzene, its single-ring analog. Their combustion

mechanisms must be similar to some extent. The initiation reactions of 1-propylnaphthalene

pyrolysis are C-H bond cleavage on the propyl chain and C-C bond cleavage listed as Reaction

(1) – (6) in Table 7-2. The former type yields 3-naphthyl-1-propyl, 1-naphthyl-2-propyl and 1-

naphthyl-1-propyl radicals, and the latter type causes the formation of 2-naphthyl-1-ethyl, 1-

naphthylmethyl, and 1-naphthyl radicals. The heats of formation of involved radicals and

molecules at 25ºC are estimated by WinMopac 7.21, a semi-empirical quantum mechanics

package, and employed to calculate the heat of reaction listed in Table 7-2. After a radical pool

of hydrogen is established, the hydrogen abstractions are listed as Reaction (7) to (11) and

replacement of the side chain by hydrogen play major roles in decomposition of the reactant

because of their much lower activation energy69 and lower heat of reactions which are

demonstrated in Table 7-257. It has been pointed out that radicals like CH3 are unimportant or not

involved in hydrogen abstraction because of the relatively large activation energy of those

reactions in the oxidation of 1-propylbenzene99. Therefore, the alkyl radicals are not considered

in hydrogen abstraction reactions in this study. It was demonstrated that the benzylic-position-

related reactions, Reaction (3), (5), (9) and (11) have the lowest heats of reaction in the same

category and are favored thermodynamically. On the other hand, the formation of 3-naphthyl-1-

105
propyl radical, Reaction (1) and (7) is the least favored because of the highest heat of reaction.

That may be the reason why 1-allylnaphthalene is a minor product.

Table 7-2 Initiation reactions and hydrogen abstraction reactions and their calculated heat of
reaction in 1-propylnaphthalene pyrolysis.

No. Reaction Heat of reaction


(kcal/mol)
CH3 CH2

1 83.53
+ H

CH3 CH3
CH
2 53.76
+ H

CH3 CH3

HC
3 52.10
+ H

CH3
CH2
4 64.43
+ CH 3

CH3

CH2
5 45.59
+ C2H5

CH3

6 C 73.38
+ C 3 H7

CH3 CH2

7 -11.53
+ H + H2

106
CH3 CH3
CH
8 -17.96
+ H + H2

CH3 CH3

HC
9 + H2 -27.75
+ H

CH3
CH2

10 -28.64
+ H + CH4

CH3

CH2
11 -39.17
+ H + C2H6

CH3

12 -24.62
+ H
+ C3H7

As shown in Figure 7-3, 1-vinylnaphthalene is the most abundant product in our experiments.

There are two important radicals coming from hydrogen abstraction of 1-propylnaphthalene

involved in 1-vinylnaphthalene formation, 1-naphthyl-1-propyl radical, followed by further β-

scission shown in Reaction (13), and 1-naphthyl-2-propyl radical. In the study of 1-

propylbenzene oxidation, Litzinger pointed out that dehydrogenation of 1-phenyl-2-propyl

radical is not significant because its final product was not observed.99 Alternatively, its

isomerization caused by 1-2 shift of the phenyl radical was proposed and styrene was formed by

further scission reaction. It was also proposed that the isomerization caused by 1-2 hydrogen

migration and forming 1-phenyl-1-propyl radical is slower than its competing reaction of

hydrogen elimination and thus unimportant99. The discussion is entirely applicable for 1-

107
naphthyl-2-propyl radical: 1-2 shift of the naphthyl radical is its major consumption pathway and

contributes to the 1-vinylnaphthalene formation by further scission shown as Reaction (14). The

1-naphthylethyl radical formed through homolysis also contributes to the formation of 1-

vinylnaphthalene via dehydrogenation, which was indicated in our ethylnaphthalene pyrolysis

study95. This pathway is supported by the analogous pathway of styrene formation in

propylbenzene pyrolysis100. The further dehydrogenation or hydrogen abstraction of the 1-

naphthyl-2-propyl radical and 3-naphthyl-1-propyl radical is responsible for the formation of

minor products, 1-allylnaphthalene and 1-isoallylnaphthalene. Another unrecognized C13H12 may

be formed by isomerization of those two compounds. In addition, β scission of 3-naphthyl-1-

propyl radical is a significant source of 1-methylnaphthalene shown in Reaction (15).

CH3
CH2
HC

+ CH3

(13)

CH3
H 3C CH2 CH2
CH

+ CH3

(14)

CH2

CH2

+ C2 H 4

(15)

Both 1, 2-di-1-naphthylethane formed through Reaction (16) and 1-methylnaphthalene, the

second most abundant products, represent the existence of a large quantity of naphthylmethyl

108
radical. While at lower temperatures, the concentration of hydrogen may be low since C-C

cleavage causing naphthylmethyl formation has relatively lower heat of reaction compared with

C-H cleavage and facilitate the formation of 1, 2-di-1-naphthylethane. When reaction

temperature increases, more hydrogen is formed and combines with naphthylmethyl radical

giving rise to methylnaphthalene. When the temperature reaches 700ºC or higher, the quantity of

naphthyl radical from aryl-alkyl C-C bond cleavage becomes important enough and forms 1, 2’-

dinaphthylmethane with naphthylmethyl radical indicated by Reaction (17). Light species

including toluene, indene, ethynylbenzene and unrecognized C10H10 are products of naphthalene

hydrogenation and further decomposition.

CH2 CH2

(16)

CH2
C
+

(17)

The formation of 1-ethylnaphthalene, acenaphthalene and acenaphthene are related to the 2-

naphthyl-1-ethyl radical, which is also shown in our previous paper.95 The latter two are related

to dehydrogenation and cyclization which was studied theoretically for PAHs101. 1-

Naphthylmethyl radical also contributes partially to the 1-ethylnaphthalene formation.

109
Naphthalene is formed from either hydrogen replacement of the side chain or combination of

naphthyl radical and hydrogen.

Based on the above description and former studies, the hypothesized scheme of reactions in 1-

propylnaphthalene pyrolysis, where major products and parts of minor products are included, is

shown in Figure 7-5. Hydrogen abstractions of 1-propylnaphthalene are not included to simplify

the scheme.

- C2 H 5
CH3

-H H2C

CH3 CH2
CH CH3
-H -H -H
1PN isomerization CH3

3-naphthyl-1-propyl

1-naphthyl-2-propyl 1allylN 1MN 2MN


- C2 H 4
isomerization
+H

H 2C CH CH2
-H CH 3 CH2
H2C CH3
C
-H 2VN + naphthyl

2-naphthyl-1-propyl
1-naphthylmethyl
1isoallylN - CH3 isomerization

CH2 dimerization 12DNM


CH3

- CH3 HC - CH3

1VN - CH3

1-naphthyl-1-propyl
-H
CH3
CH2
+H
+H- C 3 H7 -H 12DNE

1EN
1-naphthyl-2-ethyl

-H

+ H- C 2 H4

acenaphthene

- 2H

- 2H + C2H2

naphthalene
AceN

further reaction

Figure 7-5 Hypothesized scheme of reactions happening in 1-propylnaphthalene pyrolysis.


Compounds in frames are identified products and others are radicals.

110
7.3.3 Kinetic analysis

In our former study94 on the kinetic analysis of methylnaphthalene pyrolysis, it was postulated

that the consumption of reactant could be described as:


E
− a
− Ae
X = X 0e
RT t
(1)

The equation is obtained from the Arrhenius equation and plug flow assumption. X0 is the initial

quantity or concentration of reactant and is assumed as 100% in our study. A is the pre-

exponential factor (s-1), Ea is the activation energy (kJ·mol-1). R is the universal gas constant, and

T is the temperature (K).

Using a series reaction model, the formation of products or intermediates could be described

with the consideration of product decomposition as:

k1 X 0* − k1t
XP = [e − e − k 2 t ] (2)
k 2 − k1
Ea1 Ea2
Here k1 = A1e RT and k 2 = A2 e RT . k2 is the rate constant of product decomposition and could be

obtained from its pyrolysis experiments. X 0* is the theoretical initial concentration of the reactant

which will be converted to the product entirely and corresponds to Xf in Ledesma’s study20. The

estimation of all five parameters can be finished in one fitting if a large enough data pool is

available, which was impossible in our study. However, kinetic parameters of

methylnaphthalenes and 1-ethylnaphthalene pyrolysis were obtained in our previous studies94,

111
95
.Therefore, X 0* and k1 kinetic parameters of those compounds could be estimated relatively

reliably by nonlinear fitting of experimental data to the model with known k2 parameters.

Since equation (2) describes the yield of intermediates, the retention time when the yield reaches

the maximum value at temperature T could also be calculated through

∂X P
=0 (3)
∂t

The solution is:

Ea1 − Ea 2
ln( A1 / A2 ) +
t= RT (4)
Ea1 Ea2
A1e RT
− A2 e RT

And the corresponding maximum yield could be obtained by substituting t to equation (2).

The temperature where the yield of products reaches the highest value for a given reaction time

∂X P
can be calculated by = 0 . Because the final expression is too complex, it will not be
∂T

discussed in this chapter.

The experimental data in this study are fitted to equation (1) and (2), and the estimated kinetic

parameters are shown in Table 7-3. Methylnaphthalenes and 1-ethylnaphthalene are used as

examples of products because the kinetics of their pyrolysis has been studied earlier.

112
Table 7-3 Kinetic Parameters for reactant decomposition and product formation in pyrolysis of 1-
propylnaphthalene

Ea1 Ea2
lnA1 X 0* (%) lnA2
(kJ/mol) (kJ/mol)
1-propylnaphthalene 20.32 155.20
1-ethylnaphthalene 22.93 183.18 15.18 15.17 125.89
1-methylnaphthalene 33.35 268.13 15.53 22.92 224.09
2-methylnaphthalene 20.71 174.20 0.80 26.6 262.57

Figure 7-6 displays experimental data with 1 and 1.9 second retention times, and the model

simulation for the fraction of residual 1-propylnaphthalene (line). The calculation was finished at

10 K intervals in the experimental temperature range. The agreement between experimental

results and the simulation is reasonable.

120
1 sec
100 1 sec, simulated
1.9 sec
Yield (% carbon input)

1.9 sec, simulated


80

60

40

20

0
550 600 650 700 750 800
Temperature (deg C)

Figure 7-6 Model predictions and experimental results for reactant decay in 1-propylnaphthalene
pyrolysis.

Figure 7-7 shows the experimental and simulated data for methylnaphthalenes and 1-

ethylnaphthalene formation in 1-propylnaphthalene pyrolysis. The simulation is in good


113
agreement with experimental data and successfully predicts the change trend. For 1-

methylnaphthalene formation, the simulation shows that the yield is same at 940°C with 1

second and 1.9 second retention time. More yields are obtained with 1.9 second at lower

temperatures. At higher temperatures, more yields are generated with 1 second. The trend is in

agreement with the observation in the study of PAH formation from catechol pyrolysis.98 It

motivates us to know:

(1) the theoretical maximum yields of those compounds at a specific temperature without the

limitation of the retention time, and

(2) the retention time to reach those maximum yields.

The information could be used to optimize the reaction condition to decrease PAH emission or

increase the yields of some products in the chemical industry.

114
20
1MN (1 sec)
18 Simulation (1 sec)
1MN (1.9 sec)
16 Simulation (1.9 sec)

Yield (% carbon input)


14
12
10
8
6
4
2
0
550 650 750 850 950

1 Temperature (deg C)
2MN (1 sec)
Simulation (1 sec)
2MN (1.9 sec)
0.8 Simulation (1.9 sec)
Yield (% carbon input)

0.6

0.4

0.2

0
550 650 750 850 950
Temperature (deg C)
9
1EN (1 sec)
8 Simulation (1 sec)
1EN (1.9 sec)
7 Simulation (1.9 sec)
Yield (% carbon input)

0
550 650 750 850 950
Temperature (deg C)

Figure 7-7 Model predictions and experimental results for product formation in 1-
propylnaphthalene pyrolysis.

The retention time when one product’s yield reaches the highest value differs at different

reaction temperatures and can be calculated with estimated kinetic parameters by equation (4).

Furthermore, the maximum yield of the product could be obtained by equation (2) with the
115
determined retention time. Figure 7-8 demonstrated the calculated maximum yields of 1- and 2-

methylnaphthalenes and 1-ethylnaphthalene (top panel) and the retention time taken to reach

those yields (bottom panel) as functions of temperatures. The maximum yields of 1-

methylnaphthalene and 1-ethylnaphthalene are relatively stable at the studied temperature range,

while that of 2-methylnaphthalene increase from 4% to around 11%. The change trend of the

maximum yield differs for different products. With higher temperatures, the yield is raised for 1-

methylnaphthalene and 1-ethylnaphthalene, and drops slightly for 2-methylnaphthalene. The

bottom panel shows that the retention time when the yield achieves the maximum value

decreases dramatically with the increasing temperature. For example, the time for 1-

methylnaphthalene decreases from 1393 seconds at 550ºC to 2.1 millisecond at 1000ºC. Given a

retention time, the temperature at which the maximum yield is achieved and the yield can be

obtained readily. For example, when the retention time is 1 second, the temperature is 745ºC for

1-methylnaphthalene, 810ºC for 2-methylnaphthalene and 710ºC for 1-ethylnaphthalene, with

the yields of 15.1%, 0.74% and 7.65% respectively. It is consistent with the profiles shown in

Figure 7-7.

116
18

16

Maximum yield (% carbon input)


14

12

10

8
1EN
6 2MN
4 1MN

0
550 600 650 700 750 800 850 900 950 1000
Temperature (deg C)
10000

1000 1EN
2MN
100
1MN
Time (second)

10

0.1

0.01

0.001

0.0001
550 600 650 700 750 800 850 900 950 1000
Temperature (deg C)

Figure 7-8 Calculated retention time when the yields of 1- and 2-methylnaphthalene and 1-
ethylnaphthalene reach the highest values and the corresponding yields from 550 to 1000 ºC.

7.4 Conclusion

1-Vinylnaphthalene, 1, 2-di-1-naphthylethane and 1-methylnaphthalene are the most abundant

products in the pyrolysis of 1-propylnaphthalene. The initiation reactions and later hydrogen

abstractions of 1-propylnaphthalene pyrolysis were shown, and formation pathways of products

were discussed. The hypothesized scheme of reaction pathways was demonstrated. We

kinetically studied the formation of methylnaphthalenes and 1-ethylnaphthalene using a series of

reaction modeling and reactant decomposition with a first order reaction assumption. The

117
simulation results with the fitted kinetic parameters showed good agreement with the

experimental results. The retention time when the product yield reached its maximum value and

the corresponding yield as functions of the temperature were deduced from the model. The

profiles of the retention time and the maximum yield for methylnaphthalenes and 1-

ethylnaphthalene were demonstrated and the effect of temperature on that time was shown to

decrease dramatically with increasing temperature.

Reference
1. Yang, J.; Lu, M., Thermal Growth and Decomposition of Methylnaphthalenes. Environ.
Sci. Technol. 2005, 39, 3077-3082.

2. Yang, J.; Lu, M., Global Kinetic Analysis of Pyrolysis of Methylnaphthalenes. Appl.
Anal. Pyrolysis 2006.

3. Yang, J., The Study of 1-Ethylnaphthalene Pyrolysis in a Flow Reactor. Combust Flame
2006.

4. Lide, D. R.; Milne, G. W. A., Handbook of data on organic compounds. CRC Press:
Boca Raton, Fla., 1994; Vol. 4.

5. Yang, J.; Lu, M., The Preliminary Results of 1-Ethylnaphthalene Pyrolysis. In 4th Joint
Meeting of the U. S. Sections of the Combustion Institute, Philadelphia, PA, 2005.

6. Marsh, N.; Ledesma, E.; Sandrowitz, A.; Wornat, M., Yields of polycyclic aromatic
hydrocarbons from the pyrolysis of catechol [ortho-dihydroxybenzene]: Temperature and
residence time effects. Energy & fuels 2004, 18, (1), 209-217.

7. Dagaut, P.; Ristori, A.; Bakali, A. E.; Cathonnet, M., Experimental and kinetic modeling
study of the oxidation of n-propylbenzene. Fuel 2002, 81, 173-184.

8. Shaddix, C. R.; Brezinsky, K.; Glassman, I., Oxidation of 1-Methylnaphthalene. In 24th


Symp. Int. Combust., The Combustion Institute: 1992; pp 683-690.

9. Litzinger, T. A.; Brezinsky, K.; Glassman, I., Reactions of N-Propylbenzene During Gas
Phase Oxidation. Combust. and Tech. 1986, 50, 117-133.

118
10. David, A. R.; Barton;, B. D.; Stein, S. E., Thermal Decomposition of Propyl-, Isobutyl-,
and Neopentylbenzene. J. Phys. Chem. 1981, 85, 2378-2383.

11. Van Speybroeck, V.; Hemelsoet, K.; Waroquier, M.; Marin, G., Reactivity and
aromaticity of polyaromatics in radical cyclization reactions. International Journal of Quantum
Chemistry 2004 96, 568-576.

12. Ledesma, E. B.; Marsh, N. D.; Sandrowitz, A. K.; Wornat, M. J., Global Kinetic Rate
Parameters for the Formation of Polycyclic Aromatic Hydrocarbons from the Pyrolyis of
Catechol, A Model Compound Representative of Solid Fuel Moieties. Energy & Fuels 2002, 16,
1331-1336.

119
8 CHAPTER VIII CONCLUSIONS AND RECOMMENDATIONS

8.1 Conclusions

As the most abundant category of PAHs in diesel fuel, alkylated naphthalenes are closely related

to PAH and soot formation during combustion. To understand the procedure of PAH and soot

formation so that the combustion technologies could be promoted to decrease PAH and soot

emission, it is important to investigate the thermal decomposition and growth of short alkylated

naphthalenes. The experiments were conducted in a semi-isothermal tubular flow reactor at

different temperatures ranging from 550 to 1000ºC. Products including gaseous hydrocarbons,

aromatic hydrocarbons, CO, and soot were collected by the product collection system. Then the

products were identified and quantified. The trends of product yield change with reaction

temperatures were illustrated. The routes of reactant decomposition and product formation were

discussed based on experimental results and reaction mechanisms were postulated. The global

kinetic analysis method was employed to simulate reactant decomposition, and formation of

some products and results were in good agreement with experimental data.

The formation of resonance-stabilized naphthylmethyl radicals played an important role in PAH

and soot formation mechanisms in pyrolysis of methylnaphthalenes, and it provided the

120
hydrogen radical pool for the formation of naphthalene and methyl loss. The methyl radical

recombined with the reactants to form dimethylnaphthalenes, ethylnaphthalenes, and subsequent

products. The lack of carbon dimerization products in significant quantities may be due to their

continued growth into soot. The reactions by the methyl group added complexities to the

pyrolysis of methylnaphthalenes in comparison to other resonance-stabilized compounds, such as

toluene and indene. The close resemblance of the reactant consumption and product formation of

the methylnaphthalenes indicates that the same reaction mechanisms are governing the pyrolysis

of these compounds. 1-Methylnaphthalene exhibits slightly higher reactivity than 2-

methylnaphthalene.

In 2-methylnaphthalene oxidation, reactant decomposition rate was higher than that in pyrolysis

experiments resulting from the attack of oxidative species. Single ring aromatic compounds

including indene and phenylacetylene became significant products at the temperatures higher

than 850°C. Other products were similar to those in pyrolysis except that 2-naphthaldehyde was

among the products and isomerization was not apparent.

The laminar flow in the experiments of methylnaphthalene pyrolysis was characterized as a plug

flow and global kinetic analyses were conducted on the reactant decomposition. The residual

fraction of the reactant after pyrolysis starts was obtained by integration of first-order reaction

model. Experimental data were used to fit the equation and obtain kinetic parameters. Hydrogen

concentration in pyrolysis was estimated by further deduction and proved to be very low. The
121
reaction rate constants of possible decomposition routes were compared and it was shown that

the homolysis reaction became dominant at higher temperatures. Based on a series reaction

model or the calculus method and with the consideration of both product formation and further

decomposition in high temperatures, the yield of products was deducted as a function of kinetic

parameters. The model described the change trend of product yields perfectly.

1-Vinylnaphthalene, acenaphthene, 1-methylnaphthalene and naphthalene were the most

abundant products in 1-ethylnaphthalene pyrolysis. They resulted from different decomposition

pathways: hydrogen abstraction from the ethyl group at both benzylic and primary locations with

the benzylic route being favored, hydrogen displacement of the ethyl group and the hemolytic

cleavage of the ethyl group. The reactivity of 1-ethylnaphthalene is higher than that of

methylnaphthalenes, which is supported by the global kinetic analysis.

In propylnaphthalene pyrolysis, 1-vinylnaphthalene, 1,2-di-1-naphthylethane, 1-

methylnaphthalene are the most abundant products. The initiation reactions and later hydrogen

abstractions of 1-propylnaphthalene pyrolysis were shown, and formation pathways of products

were discussed. The hypothesized scheme of reaction pathways were demonstrated. We

kinetically studied the formation of methylnaphthalenes and 1-ethylnaphthalene using a series

reaction modeling and reactant decomposition with a first order reaction assumption. The

simulation result with the fitted kinetic parameters showed good agreement with experimental

results. The time when the product yield reaches its highest value as a function of temperature
122
was obtained from the model and the effect of temperature on that time for methylnaphthalenes

and 1-ethylnaphthalene was shown to decrease dramatically with increasing temperature.

8.2 Recommendation

8.2.1 Further modeling work using Chemkin

Even though reaction mechanisms of thermal decomposition of short alkylated naphthahlenes

were provided in this study qualitatively, they are very speculative. Likewise, our global kinetic

analysis was conducted based on the experimental data in a limited condition. The results are

convenient to use, but may be not accurate for other conditions. In fact, global kinetic analysis

only provides an overall prediction of decomposition and product formation, and is less able to

differentiate the various alternative pathways from a certain step, such as A to B. Some

simulation tools, such as Chemkin software, could be used to establish the model on a more

quantitative and general basis and generate more reliable results.

The Chemkin software was developed for the complex chemically reacting flow simulations with

the incorporation of complex gas phase chemical reaction mechanisms. The original developer is

102 103
Sandia National Laboratories and now Reaction Design . When specifying his problem,

one user writes an input file to declare the chemical elements in the problem including: the name

of each chemical species involved, thermochemical information about each chemical species, a

123
list of chemical reactions, and rate constant information. Bounaceur, et al. used CHEMKIN II

codes to simulate hydrocarbons pyrolysis at low temperatures and their simulation results were

in good agreement with their experimental values.104

8.2.2 Gaseous product analysis

In part of the experiments, gaseous products were collected and analyzed by different

instruments. However, not all of the products were completely identified due to the limitation of

instruments. The information about those gaseous products could provide important clues for

mechanism construction. Therefore, GC/MS is a good instrument for the identification of

gaseous products. An appropriate column may be selected and purchased for the aim.

Reference

1. Denissenko, M.; Pao, A.; Tang, M.; Pfeifer, G., Preferential formation of benzo[a]pyrene
adducts at lung cancer mutational hotspots in P53. Science 1996, 274, 430-432.

2. Bockhorn, H., Soot formation in combustion. Springer-Verlag: Berlin, Heidelberg and


New York, 1994; p 1-8, 165-192 and 253-274.

3. Mastral, A.; Callen, M., A review on polycyclic aromatic hydrocarbon (PAH) emissions
from energy generation. Environ. Sci. Technol. 2000, 34, 3051-3057.

4. NIKOLAOU, K.; MASCLET, P.; MOUVIER, G., SOURCES AND CHEMICAL-


REACTIVITY OF POLYNUCLEAR AROMATIC-HYDROCARBONS IN THE
ATMOSPHERE - A CRITICAL-REVIEW. SCIENCE OF THE TOTAL ENVIRONMENT 1984,
32 103-132.

124
5. Bjorseth, A.; Ramdahl, T., Handbook of polycyclic aromatic hydrocarbons: Emission
sources and recent progress in analytical chemistry. M. Dekker: New York, 1983; Vol. 2.

6. Dunbar, J. C.; Lin, C.-I.; Vergucht, I.; Wong, J.; Durant, J. L., Estimating the
contributions of mobile sources of PAH to urban air using real-time PAH monitoring. The
Science of the Total Environment 2001, 279, (1-3), 1-19.

7. Masclet, P.; Bresson, M.; Mouvier, G., PAH emitted by power stations and influence
of combustion parameters. Fuel 1987, 66, (1987), 556-562.

8. Song, C.; Hsu, C. S.; Mochida, I., Chemistry of diesel fuels. Taylor & Francis: New York,
2000.

9. Diesel fuel and exhaust emissions; World Health Organization: Geneva, 1996.

10. Liang, F. Composition and Formation Mechanism of Diesel Particulate Matter


Associated with Various Factors from A Non-road Diesel Generator. University of Cincinnati,
2006.

11. Rhead, M. M.; Pemberton, R. D., Sources of Naphthalene in Diesel Exhaust Emissions.
energy & fuels 1996, 10, 837-843.

12. Rhead, M.; Hardy, S., The sources of polycyclic aromatic compounds in diesel engine
emissions. Fuel 2003, 82, 385-393.

13. Williams, P.; Abbass, M.; Andrews, G., Diesel partculate emissions: the role of unburned
fuel. Combust Flame 1989, 75, 1-24.

14. Lombaert, K.; Le Moyne, L.; De Maleissye, J.; Amouroux, J., Experimental study of
PAH in engine soot by isotopic tracing Combust Sci Technol 2006, 178, (4), 707-728.

15. Wang, R.; Cadman, P., Soot and PAH production from spray combustion of different
hydrocarbons behind reflected shock waves Combust. Flame 1998, 112, (3), 359-370.

16. Tesner, P. A.; Shurupov, S. V., Soot Formation during Pyrolysis of Naphthalene,
Anthracene and Pyrene. Combust. Sci. Technol 1997, 126, 139-151.

17. Ciajolo, A.; D'Anna, A.; Barbella, R.; Bertoli, C., Combustion of Tetradecane and
Tetradecane/α-Methylnaphthalene in a Diesel Engine with regard to Soot and PAH Formation.
Combust. Sci. and Tech. 1992, 87, 127-137.

18. Brezinsky, K.; Hura, H.; Glassman, I., Oxidation/Pyrolysis chemistry as related to fuel
sooting tendencies. Energy & fuels 1988, 2, (4), 487-493.

125
19. Burnham, A. K.; Braun, R. L., Global Kinetic Analysis of Complex materials. energy &
fuels 1999, 13, (1), 1-22.

20. Ledesma, E. B.; Marsh, N. D.; Sandrowitz, A. K.; Wornat, M. J., Global Kinetic Rate
Parameters for the Formation of Polycyclic Aromatic Hydrocarbons from the Pyrolyis of
Catechol, A Model Compound Representative of Solid Fuel Moieties. Energy & Fuels 2002, 16,
1331-1336.

21. Mansurov, Z., Soot formation in combustion processes. Combustion, Explosion, and
Shock Waves 2005, 41, (6), 727-744.

22. Kiefer, J.; Mizerka, L. J.; Patel, M.; Wei, H., A shock-tube investigation of major
pathways in the high-temperature pyrolysis of benzene. J. Phys. Chem 1985, 89, 2013-2019.

23. Boehm, H.; Jander, H.; Tanke, D. In PAH growth and soot formation in the pyrolysis of
acetylene and benzene at high temperatures and pressures: Modeling and experiment, 12th
Symposium (International) on Combustion, Pittsburgh, 1998; The Combustion Institute:
Pittsburgh, 1998; pp 1605-1612.

24. Richter, H.; Howard, J. B., Formation of polycyclic aromatic hydrocarbons and their
growth to soot -- a review of chemical reaction pathways. Progress in Energy and Combustion
Science 2000, 26, 565-608.

25. Majidi, V.; Saito, K.; Gordon, A.; Williams, F., Laser-desorption time-of-flight mass-
spectrometry analysis of soot from various hydrocarbon fuels. Combust Sci Technol 1999, 145,
37-56.

26. Badger, G. M.; Novotny, J., Formation of aromatic hydrocarbons at high temperatures.
xii Pyrolysis of bezene. J. Chem. Soc. 1961, 3400-3402.

27. Richter, H.; Grieco, W. J.; Howard, J. B., Formation mechanism of polycyclic aromatic
hydrocarbons and fullerenes in premixed benzene flames. Combust Flame 1999, 119, 1-22.

28. Bittner, J. D.; Howard, J. B. In Composition profiles and reaction mechanisms in a near-
sooting premixed benzene/oxygen/argon flame, 18th Symposium (international) on Combustion,
1981; the combustion institute: 1981; pp 1105-1116.

29. Chin, J. S.; Lefebvre, A. H., Influence of fuel chemical properties on soot emissions from
gas turbine combustors. Combust Sci Technol 1990, 73, 479-486.

30. Badger, G. M.; Jolad, S. D.; Spotswood, T. M., The Formation of Aromatic
Hydrocarbons at High Temperatures: xx. The Pyrolysis of [1-14C]Naphthalene. Aust. J. Chem.
1964, 17, 771-777.

126
31. Griesheimer, J.; Homann, K. H. In Large molecules, radicals ions, and small soot
particles in fuel-rich hydrocarbon flames Part II. Aromatic radicals and intermediate PAHs in a
premixed low-pressure naphthalene/oxygen/argon flame, 27th Symposium (International) on
Combustion, Pittsburgh, 1998; The Combustion Institute: Pittsburgh, 1998; pp 1753-1759.

32. Sarofim, A.; Longwell, J.; Wornat, M.; Mukherjee, J., The role of biaryl reaction in PAH
and soot formation. In Soot formation in combustion: mechanisms and models, Bockhorn, H., Ed.
1994; pp 485-496.

33. Masonjones, M. C.; Lafleur, A. L.; Sarofim, F., Biarene formation during pyrolysis of a
mixture of anthracene and naphthalene. Combust Sci Technol 1995, 109, 273-285.

34. Smith, C. M.; Savage, P. E., Reactions Of Polycyclic Alkylaromatics .5. Pyrolysis Of
Methylanthracenes. Aiche Journal 1993, 39, (8), 1355-1362.

35. Smith, R. D., A direct mass spectrometric study on the mechanism of toluene pyrolysis at
high temperature. J. Phys. Chem. 1979, 83, 2553-1562.

36. Smith, R. D., Formation of radicals and complex organic compounds by high-
temperature pyrolysis: the pyrolysis of toluene. Combust Flame 1979, 35, 179-190.

37. Brooks, C. T.; Cummins, C. P. R.; Peacock, S. J., Pyrolysis of toluene using a static
system. Trans Faraday Soc. 1971, 67, 3265-3274.

38. Errede, L. A.; Cassidy, J. P., Formation of aromatic hydrocarbons at high temperatures.
IX. Pyrolysis of toluene, ethylbenzene, propylbenzene, and butylbenzene, . J. Am. Chem. Soc.
1960, 82, 3653-3658.

39. Lindstedt, R. P.; Maurice, L. Q., Detailed kinetic modeling of toluene combustion.
Combust Sci Tech 1996, 120, 119-167.

40. Emdee, J.; Brezinsky, K.; GLASSMAN, I., A Kinetic-model for the Oxidation of
Toluene near 1200 K. J. Chem. Phys. 1992, 96, 2151-2161.

41. Jess, A., Mechanisms and kinetics of thermal reactions of aromatic hydrocarbons from
pyrolysis of solid fuels. Fuel 1996, 75, 1441-1448.

42. Robaugh, D. A.; Stein, S. E., Very-low-pressure Pyrolysis of Ethylbenzene,


Isopropylbenzene, and Tert-butylbenzene. Int. J. Chem. Kinet. 1981, 13, 445 -462.

43. Davis, H. G., Rate of formation of toluene from ethylbenzene. Int. J. Chem. Kinet. 1983,
15, 469-474.

127
44. Brouwer, L.; Muller-Markgraf, W.; Troe, J. In Identification of Primary Reaction
Products in the Thermal Decomposition of Aromatic Hydrocarbons, 20th Symp. Int. Combust.,
1984; The Combustion Institute: 1984; pp 799-806.

45. Mizerka, L. J.; Kiefer, J. H., The High-temperature Pyrolysis of Ethylbenzene Evidence
for Dissociation to Benzyl and Methyl Radicals. Int. J. Chem. Kinet. 1986, 18, 363-378.

46. Pamidimukkala, K. M.; Kern, R. D., The High Temperature Pyrolysis of Ethylbenzene.
Int. J. Chem. Kinet. 1986, 18, 1341-1353.

47. Smith, C.; Savage, P., Reactions of Polycyclic Alkylaromatics:3. Struture and Reactivity.
AIChE Journal 1991, 37, 1613-1624.

48. Tancell, P.; Rhead, M.; Pemberton, R.; Braven, J., Diesel combustion of an alkylated
polycyclic aromatic hydrocarbon. FUEL 1996 75, 717-723.

49. Smith, C. M.; Savage, P. E., Reactions Of Polycyclic Alkylaromatics .4. Hydrogenolysis
Mechanisms In 1-Alkylpyrene Pyrolysis. Energy & Fuels 1992, 6, (2), 195-202.

50. Freccero, M.; Golfi, R.; Sarzi, A. M.; Rastelli, A., Origin of Pyrene under High
Temperature Conditions in the Gas Phase. The Pivotal Role of Phenanthrene. J. Org. Chem 1999,
64, 3861-3866.

51. Marsh, N. D.; Wornat, M. J. In Formation pathways of ethhynyl-subsituted and


cycloentafused polycyclic aromatic hydrocarbons, Proceedings of Comubsiton Institute, 2000;
2000; pp 2585-2592.

52. Wornat, M. J.; Ledesma, E. B., C16H10 etynyl-substituted polycyclic aromatic


hydrocarbons from the pyrolysis of coal, coal volatiles, and anthracene. Polycylic Aromatic
Compounds 2000, 18, 129-147.

53. Wornat, M. J.; Vriesendorp, F. J. J.; Lafleur, A. L.; Plummer, E. F.; Necula, A.; Scott, L.
T., The identification of new ethynyl-substituted and cyclopenta-fused polycyclic aromatic
hydrocarbons in the products of anthracene pyrolysis. Polycylic aromatic compounds 1999, 13,
221-240.

54. Marsh, N. D.; Wornat, M. J.; Scott, L. T.; Necula, A.; Lafleur, A. L.; Plummer, E. F., The
identification of cyclopenta-fused and ethynyl-subsitituted polycyclic aromatic hydrocarbons in
benzene droplet combustion products. Polycylic aromatic compounds 2000, 13, 379-402.

55. Behar, F.; Budzinski, H.; Vandenbroucke, M.; Tang, Y., Methane Generation from Oil
Cracking: Kinetics of 9-Methylphenanthrene Cracking and Comparison with Other Pure
Compounds and Oil Fractions. Energy & fuels 1999, 13, 471-481.

128
56. Lorant, F.; Behar, F.; Vandenbroucke, M.; Mckinney, D. E.; Tang, Y., Methane
generation from methylated aromatics: kinetic study and carbon isotope modeling, Energy Fuels.
Energy & fuels 2000, 14, 1143-1155.

57. Shaddix, C. R.; Brezinsky, K.; Glassman, I. In Oxidation of 1-Methylnaphthalene, 24th


Symp. Int. Combust., 1992; The Combustion Institute: 1992; pp 683-690.

58. Shaddix, C. R.; Brezinsky, K.; Glassman, I., Analysis of Fuel Decay Routes in the High-
Temperature Oxidation of 1-Methylnaphthalene. Combust. Flame 1997, 108, 139-157.

59. Ledesma, E. B.; Kalish, M. A.; Nelson, P. F.; Wornat, J. J.; Mackie, J. C., Formation and
fate of PAH during the pyrolysis and fuel-rich combustion of coal primary tar. Fuel 2000, 79,
1801-1814.

60. Vovelle, C.; Delfau, J. L., Formation of aromatics in combustion systems. In Pollutants
from combustion : formation and impact on atmospheric chemistry, Vovelle, C., Ed. Kluwer
Academic Publishers: Boston, 2000.

61. NIST Chemistry WebBook, NIST Standard Reference Database Number 69, June 2005
Release. In 2005.

62. Weast, R. C.; Grasselli, J. G., CRC Handbook of Data on Organic Compounds. second
ed.; CRC Press, Inc.: Boca Raton, FL, 1989; Vol. 1.

63. Eichelberger, J. W.; Munch, J. W.; Shoemaker, J. A., EPA 525.2 Determination of
organic compounds in drinking water by liquid-solid extraction and capillary column gas
chromatography/mass spectrometry. In Varian Environmental Application Manual.

64. El-kady, A. M.; Jeng, S.-M., Experimental Investigation of Temperature and Species
Concentration Characteristics of Swirling Spray Combustion. In 43rd Aerospace Sciences
Meeting & Exhibit, Reno, Nevada, 2005.

65. Seinfeld, J. H.; Pandis, S. N., Atmospheric Chemistry and Physics: From Air Pollution to
Climate Change. John Wiley & Sons: New York, 1997.

66. Lu, M. Carbon growth from cyclopentadienyl moities in combustion. Georgia Institute of
Technology, Atlanta, GA, 2000.

67. Lu, M.; Mulholland, J. A., PAH Growth from the Pyrolysis of CPD, Indene and
Naphthalene Mixture. Chemosphere 2004, 55, 605-610.

68. Skjoth-Rasmussen, M.; Glarborg, P.; Ostberg, M.; Johannessen, J.; Livbjerg, H.; Jensen,
A.; Christensen, T., Formation of polycyclic aromatic hydrocarbons and soot in fuel-rich
oxidation of methane in a laminar flow reactor. Combust Flame 2004, 2004, (136), 91-128.

129
69. Dagaut, P.; Ristori, A.; Bakali, A. E.; Cathonnet, M., Experimental and kinetic modeling
study of the oxidation of n-propylbenzene. Fuel 2002, 81, 173-184.

70. Yang, J.; Lu, M., Thermal Growth and Decomposition of Methylnaphthalenes. Environ.
Sci. Technol. 2005, 39, 3077-3082.

71. Ramayya, S. V.; M.J Antal, J., Evaluation of systematic error incurred in the plug flow
idealization of tubular flow reactor data. energy & fuels 1989, 3, (1989), 105-108.

72. Cutler, A. H.; Antal, M. J.; Jones, M., A Critical Evaluation of the Plug-flow Idealization
of Tubular-flow Reactor Data. Ind. Eng. Chem. Res. 1988, 27, 691-697.

73. Furue, H.; Pacey, P. D., Performance of Cylindrical Flow Reactors in a Kinetic Study of
the Isomerization of Cyclopropane. J. Phys. Chem. 1980, 84, 3139-3143.

74. Cussler, E. L., Diffusion Mass Transfer in Fluid Systems (2nd Edition). Cambridge
University Press: Cambridge, United Kingdom, 1997.

75. Kimura, Y.; Abe, D.; Terazima, M., Vibrational Energy Relaxation of Naphthalene in the
S1 State in Various Gases. J. Chem. Phys. 2004, 121, (12), 5794-5800.

76. Yaws, C. L., Handbook of Viscosity V. 4. Inorganic Compounds and Elements. Gulf Pub.
Co.: Houston, Texus, 1995; Vol. 4.

77. Lee, J. C.; Yetter, R. A.; Dryer, F. L.; Tomboulides, A. G.; Orszag, S. A., Simulation and
Analysis of laminar flow reactors. Combust. Sci. and Tech. 2000, 159, 213-235.

78. Dang, V. D.; Steinberg, M., Convective Diffusion with Homogeneous and Heterogeneous
Reactions in a Tube. J. Phys. Chem. 1980, 84, 214-219.

79. Saxena, S. C., Devolatilization and Combustion Charateristics of Coal Particles. Prog.
Energy Combust. Sci. 1990, 16, 55-94.

80. Yang, J.; Lu, M., One Improved Global Kinetic Analysis Based on Combustion
Experiments at Different Temperatures. In 4th Joint Meeting of the U. S. Sections of the
Combustion Institute, Philadelphia, PA, 2005.

81. Berthouex, P. M.; Brown, L. C., Statistics for Environmental Engineers (2nd Edition).
Lewis Publishers: Boca Raton, Florida, 2002; p 397.

82. Lapin, L. L., Probability and Statistics for Modern Engineering (2nd Edition). PWS-
KENT: Boston, 1990.

130
83. Cvetanovic, R. J.; Singlelon, D. L.; Paraskevopoulos, G., Evaluations of the Mean Values
and Standard Errors of Rate Constants and Their Temperature Coefficients. J. Phys. Chem. 1979,
83, (1), 50-60.

84. Pitsch, H., Detailed Kinetic Reaction Mechanism for Ignition and Oxidation of alfa-
Methylnaphthalene. In 26th Symp. Int. Combust., The Combustion Institute: 1996; pp 721-728.

85. Fogler, H. S., Elements of Chemical Reaction Engineering. Prentice-Hall: Englewood


Cliffs, New Jersey, 1986.

86. Mimura, K.; Madono, T.; Toyama, S.; Sugitani, K.; Sugisaki, R.; Iwamatsu, S.; Murata,
S., Shock-induced Pyrolysis of Naphthalene and Related Polycyclic Aromatic Hydrocarbons
(Anthracene, Pyrene, and Fluoranthene) at Pressures of 12-33.7 GPa. J. Anal. Appl. Pyrolysis
2004, 72, 273-278.

87. Platonov, V. V.; Ivleva, L. N.; Vol-Epshtein, A. B.; Kasimtseva, T. V.; Klyavina, O. A.,
Kinetics of the Homogenous Pyrolysis of Naphthalene and Monomethylnaphthalenes. Solid Fuel
Chemistry (English Translation of Khimiya Tverdogo Topliva) 1986, 20, 74-82.

88. Lijinsky, W.; Raha, C. R., The Pyrolysis of 2-Methylnaphthalene. J. Org. Chem 1961, 26,
3566-3568.

89. Beltrame, P.; Carnltl, P.; Maronglu, B.; Mura, L.; Solinas, V.; Mori, S.,
Hydrodealkylation of 2-Ethylnaphthalene. A Study of Reaction Products and Kinetics. Ind. Eng.
Chem. Process Des. Dev. 1979, 18, 338-342.

90. Beltrame, P.; Carnltl, P.; Maronglu, B.; Mura, L.; Solinas, V.; Mori, S.,
Hydrodealkylation of 1-Ethylnaphthalene and its Side Reactions. A Kinetic Study. Ind. Eng.
Chem. Process Des. Dev. 1981, 20, 379-384.

91. Curran, H.; C.Wu; Marinov, N.; Pitz, W. J.; Westbrook, C. K.; Burcat, A., The Ideal Gas
Thermo-dynamics of Diesel Fuel Ingredients. I. Naphthalene Derivatives and their Radicals. J.
Phys. Chem. Ref. Data 2000, 29, 463-517.

92. McMillen, D. F.; Golden, D. M., Hydrocarbon Bond Dissociation Energies Ann. Rev.
Phys. Chem. 1982, 33, 493-532.

93. Litzinger, T. A.; Brezinsky, K.; Glassman, I., The Oxidation of Ethylbenzene near 1060
K. Combust Flame 1986, 63, 251-267.

94. Yang, J.; Lu, M., Global Kinetic Analysis of Pyrolysis of Methylnaphthalenes. Appl.
Anal. Pyrolysis 2006.

131
95. Yang, J., The Study of 1-Ethylnaphthalene Pyrolysis in a Flow Reactor. Combust Flame
2006.

96. Lide, D. R.; Milne, G. W. A., Handbook of data on organic compounds. CRC Press:
Boca Raton, Fla., 1994; Vol. 4.

97. Yang, J.; Lu, M. In The Preliminary Results of 1-Ethylnaphthalene Pyrolysis, 4th Joint
Meeting of the U. S. Sections of the Combustion Institute, Philadelphia, PA, March 20th-23rd,
2005, 2005; Philadelphia, PA, 2005.

98. Marsh, N.; Ledesma, E.; Sandrowitz, A.; Wornat, M., Yields of polycyclic aromatic
hydrocarbons from the pyrolysis of catechol [ortho-dihydroxybenzene]: Temperature and
residence time effects. Energy & fuels 2004, 18, (1), 209-217.

99. Litzinger, T. A.; Brezinsky, K.; Glassman, I., Reactions of N-Propylbenzene During Gas
Phase Oxidation. Combust. and Tech. 1986, 50, 117-133.

100. David, A. R.; Barton;, B. D.; Stein, S. E., Thermal Decomposition of Propyl-, Isobutyl-,
and Neopentylbenzene. J. Phys. Chem. 1981, 85, 2378-2383.

101. Van Speybroeck, V.; Hemelsoet, K.; Waroquier, M.; Marin, G., Reactivity and
aromaticity of polyaromatics in radical cyclization reactions. International Journal of Quantum
Chemistry 2004 96, 568-576.

102. http://www.ca.sandia.gov/chemkin/index.html. (Dec. 2006),

103. http://www.reactiondesign.com/lobby/open/index.html. (Dec. 2006),

104. Bounaceur, R.; Warth, V.; Marquaire, P.; Scacchi, G.; Domine, F.; Dessort, D.; Pradier,
B.; Brevert, O., Modeling of hydrocarbons pyrolysis at low temperature. Automatic generation
of free radicals mechanisms. Journal of analytical and applied pyrolysis 2002, 64, 103-122.

132
9 APPENDIX A CALIBRATION CURVES FOR THE SYRINGE PUMP AND THE
FLOW METER
nomnial speed vs real flow rate
18

16 x1 x1/10

14

12
flow rate (ul/min)

10

0
0 5 10 15 20 25 30 35 40 45 50 55
Reading

Figure A-1 The calibration curve for the syringe pump

133
Calibration of Flow Meter
6

5.5

4.5

4
flow rate, lpm

3.5

2.5

2
in system
1.5 Real data in 1 atm
Provided data
1 In system+nozzle
0.5

0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90 95 100 105
reading

Figure A-2 Calibration curves for the flow meter in Helium. “In system” represents that the calibration condition is same as in practical
experiments except high temperatures. “Real data in 1 atm” represents that the flow meter is connected to the bubble meter directly.
“Provided data” means the data provided by the manufacturer. “In system+nozzle” condition is similar to that of “In system” except that
one noozle is connected to the bubble meter.

134
3.0

2.8

2.6

Flow rate (lpm) 2.4

2.2

2.0

1.8

1.6

1.4

1.2

1.0
30 40 50 60 70 80 90 100 110 120 130 140 150 160
Reading

Figure A-3 The calibration curve for the flow meter in Nitrogen

135
10 APPENDIX B EXPERIMENTAL DATA OF AVERAGE PRODUCT YIELDS

Table B-1 Product yields in 1-methylnapthalene pyrolysis in % carbon input

Molecular
Name 800 850 900 925 950 975 1000
Structure
C10H8(2) 2.364731 5.25189 14.9447 21.68596 27.12005 21.67536 21.78058
C11H10 2-Methylnaphthalene 0.780706 3.333259 10.99713 10.78081 7.050999 0.76752 0.420397
C12H10(4) 0.728644 2.250878 1.842898 1.478606 0.697043 0.065714 0.004254
C12H12(9) 0.512181 0.497755 0.576353 0.387504 0.096334 0.121688 0.000001
C12H8 Acenaphthalene 0.103035 0.384327 1.512202 2.335348 2.969343 2.538251 2.607904
C13H10(7) 0.055722 0.354426 0.461556 0.576819 0.621527 0.100679 0.103601
C14H10(2) 0.067561 0.159609 0.416954 0.60083 0.793575 0.328081 0.451789
C16H10(3) 0 0 0 0.054261 0.108917 0.043226 0.043226
C20H14(3) 0.0001 0.00001 0.284143 0.634842 1.311739 1.469614 0.557115
C6H6 Benzene 0.091485 0.039606 0.098868 0.120975 0.190393 0.454073 0.454073
C7H8 Toluene 0.032836 0.042127 0.074823 0.065206 0.061911 0.041004 0.049739
C8H10 m-Xylene 0.049004 0.024982 0.031657 0 0 0 0
C8H8 Styrene 0.031056 0.026186 0.085928 0.019605 0.018202 0 0
C9H8 Indene 0.069949 0.108216 0.285176 0.269992 0.272874 0.198752 0.16384
C11H10 1-Methylnaphthalene 84.86479 71.21102 42.39682 16.24077 7.182699 5.694239 0.282716

136
Table B-2 Product yields in 2-methylnapthalene pyrolysis in % carbon input

Molecular
Name 800 850 900 925 950 975 1000
Structure
C10H8(2) 0.11 2.2660242 6.8065379 14.36973 19.53832 20.00669 20.75478

C10H10 1-Methylnaphthalene 0.64 2.1334871 5.5034226 5.95997 4.036824 0.677647 0.210341

C12H10(4) 0.039 1.6564481 1.0111334 0.865003 0.582759 0.847985 0.049784

C12H12(8) 0.038 0.248722 0.4373771 0.240904 0.14721 0.00443 0.00001

C12H8 Naphthalene 0.012 0.1189805 0.5295163 1.257143 1.874554 1.463516 3.13984

C13H10(7) 0.074 0.10487 0.1647583 0.383744 0.448786 0.117755 0.028144

C14H10(2) 0.018 0.126232 0.0842849 0.223367 0.319786 0.223178 0.065954

C16H10(3) 0 0 0 0.016642 0.03311 0.047239 0.053629

C20H14(3) 0.000001 0.00001 0.1578068 0.658618 0.973974 0.924426 0.651209

C7H8 Toluene 0.007 0.0553244 0.0443397 0.200041 0.057932 0.220451 0.063803

C9H8 Indene 0.004 0.1176884 0.0764934 0.130747 0.187293 0.26589 0.188984

C8H8 Styrene 0.0049985 0.0722658 0.08543 0.083141 0.094047 0.011058

C12H10 2-Ethylnaphthalene 0.038497 0.1752496 0.2671324 0.106325 0.051148 0.00443 0

137
Table B-3 Product yields in 2-methylnapthalene oxidation % carbon input

Molecular
Name 650 700 750 800 850 900 950
Structure
C6H6 Benzene 0.003534 0.003426 0.01831 0.083646 1.083751 1.302657 2.935693
C7H8 Toluene 0.003132 0.004018 1.021919 1.230559 0.407584 0.988348 2.145116
C8H10 Ethylbenzene 0 0 0.021626 0.032002 0.01965 0.023986 0.018615
C8H10 m-Xylene 0 0 0.074873 0.095109 0.025109 0.061164 0.111241
C8H6 Phenylethylene 0 0 0.02607 0.225119 2.109411 2.436627 3.260265
C8H8 Stylene 0.00241 0.007552 0.053007 0.082643 0.466584 0.505548 0.718062
C7H6O Benzaldehyde 0 0 0.014425 0.032607 0.151881 0.08506 0.048268
C9H8 0 0 0.029201 0.058144 0.064071 0.320685 0.075195
C9H8 0 0 0.021495 0.043618 0.037177 0.260621 0.054972
C8H6O Benzofuran 0 0 0 0.013237 0.126833 0.094783 0.050564
C9H8 Indene 0.007315 0.01025 0.066426 0.503101 2.726376 3.496175 3.883044
C8H8O m-Tolualdehyde 0 0 0 0.006242 0 0 0
C10H10 1H-Indene, methyl 0 0 0.019086 0.066677 0.156652 0.208252 0.037795
C10H10 Naphthalene,1-2dihydro- 0 0 0.014063 0.542524 0.099332 0.117761 0.023991
C10H8 Benzofulvene 0 0 0.012654 0.069218 0.226878 0.262797 0.08427
C10H8 Naphthalene 0.082028 0.118631 0.357744 1.162379 8.498102 9.914233 16.92932
C11H10 2-Methylnaphthalene 89.24297 80.91613 74.96698 63.53221 15.4706 11.45922 2.870383
C11H10 1-Methylnaphthalene 0.727476 0.683218 0.595761 0.662585 1.131722 1.211205 0.700962

138
C12H10 Biphenol 0 0 0.000828 0 0.008641 0.032273 0.019763
C12H12 2-Eethylnaphthalene 0.019726 0.039259 0.21707 0.428429 0.07455 0.046232 0.00734
C12H8 Biphenolyene 0 0 0 0.070041 0.13073
C12H10 1-Vinylnaphthalene 0 0 0 0 0.10858 0.129554 0.048579
C12H12 1,7-Dimethylnaphthalene 0 0 0 0.012709 0.041169 0.028722 0.008036
C12H12 1,6-Dimethylnaphthalene 0 0.001855 0.003269 0.004165 0.044507 0.022334 0.009222
C12H10 2-Vinylnaphthalene 0 0 0 0.039894 0.215282 0.159368 0.063022
C12H12 1,4-Dimethylnaphthalene 0.000927 0 0.001461 0.004565 0.014836 0.011225 0.006877
C12H12 1,5-Dimethylnaphthalene 0 0 0 0.003586 0.014836 0.006706 0
C12H8 Acenaphthylene 0 0 0 0.004413 0.271787 0.402261 1.809398
C10H8O 1-Naphthol 0 0 0 0 0 0.114855 0.022998
C11H8O 1-Naphthaldehyde 0 0 0 0 0 0.141505 0.037622
C11H8O 2-Naphthaldehyde 0 0.224658 0.448472 0.718259 0.910142 0.27951 0.020003
C13H10 Fluorene 0 0 0 0.001074 0.165221 0.311215 0.716794
C13H10s 0 0 0 0 0.228902 0.298799 0.19035
C14H10 Phenanthrene 0 0 0 0 0.735703 0.183297 0.740796
C14H10 Anthracene 0 0 0 0 0.172928 0.18083 0.18584

139
Table B-4 Product yields in 1-eethylnapthalene pyrolysis in % carbon input

Molecular
Name 600 650 700 750 800 850
Structure
C10H8 Naphthalene 0.81001 1.776554 2.860655 4.014783 4.069198 8.710909
C11H8 1-Methylnaphthalene 0.190484 0.923264 3.97246 12.19252 21.05531 29.85717
C11H8 2-Methylnaphthalene 0 0.005823 0.051948 0.216782 1.066538 5.540476
C12H10 2-Vinylnaphthalene 0 0.259571 0.202852 0.447943 1.002451 2.254261
C12H10 1-Vinylnaphthalene 7.674534 18.04656 42.55106 54.82157 50.51066 38.51282
C12H10 Acenaphthene 0.919301 1.925502 4.16073 3.703032 2.316069 1.402199
C12H12 1,7-Dimethylnaphthalene 0 0 0.011497 0.036027 0.223058 0.360873
C12H12 1,6-Dimethylnaphthalene 0 0 0.002858 0.003665 0.111877 0.029656
C12H12 1-Ethylnaphthalene 91.18565 77.50758 46.92368 23.04893 8.517619 0.258929
C12H12 1,4-Dimethylnaphthalene 0 0 0.008602 0.023515 0.141926 0.099996
C12H12 1,5-Dimethylnaphthalene 0 0.002276 0.008342 0.00672 0.02662 0
C12H10 Acenaphthalene 0.023158 0.136987 1.185672 3.476582 4.871322 9.911577
C13H10 Fluorene 0 0 0 0.002356 0 0.05454
C14H10 Phenanthrene 0 0 0 0 0.000787 0.046479
C9H8 Indene 0 0 0.003772 0.028086 0.056542 0.087324

140
Table B-5 Product yields in 1-propylnapthalene pyrolysis in % carbon input for 1 second

Molecular Name 550 600 650 700 750


Structure

C7H8 Toluene 0 0.004012 0.001246 0.00282 0.029418


C8H6 Phenylethyne 0 0 0.005322 0.029126 0.013994
C9H8 Indene 0.000941 0.00075 0.005549 0.056619 0.064449
C10H10 2-Methyl-1H-indene 0 0 0.00526 0.00832 0.008561
C10H8 Naphthalene 0.063723 0.364224 1.849494 4.142566 4.378041
C11H10 2-Methylnaphthalene 0.011978 0.008424 0.093678 0.345882 0.542036
C11H10 1-Methylnaphthalene 0.065648 0.348306 3.075507 11.84593 14.74734
C12H12 1-Ethylnaphthalene 0.104544 0.727827 4.347437 7.024056 5.07719
C12H10 1-Vinylnaphthalene 1.930585 10.31101 32.3026 36.64516 34.78847
C12H12 1,7-Dimethylnaphthalene 0.002099 0 0.001016 0.009727 0.013691
C12H10 2-Vinylnaphthalene 0.042761 0.041493 0.165635 0.374039 0.372149
C12H12 1,4-Dimethylnaphthalene 0.002565 0.003076 0.00269 0.006249 0.007486
C12H8 Acenaphthalene 0.019149 0.010461 0.39795 2.396649 2.977773
C12H12 1,2-Dimethylnaphthalene 0.002332 0.001927 0.004104 0.011156 0.012501
C13H12 1-Allylnaphthalene 0.003285 0.005032 0.018798 0.041075 0.028821
C12H10 Acenaphthene 0 0.028 0.271 0.484 0.343
C13H14 1-n-Propylnaphthalene 97.87013 64.59233 30.70152 2.070442 0.033773
C13H10 Fluorene 0 0 0.002103 0.005337 0.008698
C13H12 1-Isoallylnaphthalene 0.003121 0.012753 0.099216 0.156883 0.193602
(17.896)
C14H10 Phenanthrene 0 0 0.005194 0.007365 0.01578
C17H14 Naphthalene, 2- 0 0.001131 0.007463 0.009429 0.007527
(phenylmethyl)-
C20H14 1,1-Binaphthyl 0 0.002508 0.009372 0.010088 0.007771
C20H14 2,2-Binaphthyl 0 0 0.00208 0.001907 0.000139
C21H16 1,1'-Methylenebis- 0 0 1E-08 0.552566 1.272808
naphthalene
C22H18 1,1'-(1,2-Ethanediyl)bis- 0.165057 1.910991 15.21571 10.38682 2.346342
naphthalene

141
Table B-6 Product yields in 1-propylnapthalene pyrolysis in % carbon input for 1.9 second

Molecular
Name 550 600 650 700 750
Structure
C7H8 Toluene 0 0.000455 0.00217 0.002 0.004179
C8H6 Phenylethyne 0 0.000944 0.006561 0.018 0.018524
C9H8 Indene 0 0.002335 0.009608 0.035 0.075427
C10H10 2-Methyl-1H-indene 0 0.005364 0.008863 0.011 0.011407
C10H8 Naphthalene 0.038704 0.728894 3.295405 5.433 7.492797
C11H10 2-Methylnaphthalene 0 0.067036 0.149926 0.427 0.750024
C11H10 1-Methylnaphthalene 0.026874 0.696796 4.790252 11.704 17.76413
C12H12 1-Ethylnaphthalene 0.115171 1.840854 7.022769 6.155 2.539105
C12H10 1-Vinylnaphthalene 2.865212 20.97361 48.88314 46.812 38.59435
C12H13 1,7-Dimethylnaphthalene 0.006337 0.006997 0.003139 0.006 0.029997
C12H10 2-Vinylnaphthalene 0.00494 0.072581 0.346397 0.498 0.725882
C12H14 1,4-Dimethylnaphthalene 0.004014 0.002315 0.00489 0.008 0.017745
C12H8 Acenaphthene 1.566247 0.064888 1.374898 6.404 7.173007
C12H15 1,2-Dimethylnaphthalene 0.005704 0.002315 0.008 0.013 0.020702
C13H12 1-Allylnaphthalene 0.016788 0.007456 0.034343 0.041 0.026351
C12H16 1,8-Dimethylnaphthalene 0 0 0 0.000 0.005915
C12H10 Acenaphthene 0 0.126 0.573 0.761 0.335
C13H14 1-n-Propylnaphthalene 80.94688 53.62879 21.29652 3.761 0.155885
C13H10 Fluorene 0 0 0 0.011 0.013764
C14H10 Phenanthrene 0 0.006815 0.006505 0.008 0.048383
C17H14 Naphthalene, 2- 0 0.001864 0.001779 0.005 0.013278
(phenylmethyl)-
C20H14 1,1'-Binaphthyl 0 0.004235 0.004042 0.007 0.036328
C20H14 2,2'-Binaphthyl 0 0.004 0.003817 0.000 0
C21H16 1,1'-Methylenebis- 0 0 0 0.360 0.855897
naphthalene
C22H18 1,1'-(1,2-Ethanediyl)bis- 0.097933 3.50342 11.31879 1.213 0
naphthalene

142
11 APPDENDIX C EXPERIMENTAL DATA OF YIELDS OF

RESIDUAL REACTANTS AND SOME PRODUCTS USED

IN GLOBAL KINETIC ANALYSIS

Table C-1 1-Methylnaphthalene pyrolysis

Residual
Temperature 2-Methylnaphthalene
1-Methylnaphthalene
800 82.24 0.786348
800 78.14 0.677617
800 94.22 0.878153
850 69.4239291 4.263038
850 66.24815362 3.438802
850 77.96098399 2.297939
900 39.21713442 10.94646
900 42.72525849 11.18857
900 48.01899926 11.2622
900 39.62587232 10.59127
925 19.57163959 11.71099
925 12.9098966 9.850635
950 4.062038405 5.763861
950 6.277695716 7.30501
950 9.854763029 6.996589
950 8.799981267 8.026157
950 6.919016323 7.163377
975 5.694239291 1.246933
1000 0.184638109 0.288107
1000 0.380794511 0.552686

143
Table C-2 2-Methylnaphthalene pyrolysis

Residual
Temperature 1-Methylnaphthalene
2-Methylnaphthalene
800 0.64 96.87861617
850 2.133487087 74.15482874
900 4.131086054 40.93314093
900 6.875759159 60.32305906
925 5.750298716 25.95183115
925 6.169641161 19.92325023
950 2.635316355 7.290954659
950 5.438331639 16.2374462
975 0.478574445 1.097948667
975 0.876720015 2.066692949
1000 0.210340942 0.472445136

144
Table C-3 1-Ethylnaphthalene pyrolysis

Temp Residual
Naphthalene 1-Methylnaphthalene 2-Methylnaphthalene
(ºC) 1-Ethylnaphthalene
600 0.727605 0.176499 0 92.92328
600 0.892414 0.204469 0 89.44802
650 1.882206 1.183526 0.011645 77.15166
650 1.670902 0.663002 0 77.86349
700 3.347695 4.304389 0.067302 49.09786
700 2.373616 3.640531 0.036594 44.74949
750 4.016919 9.752345 0.275684 19.01511
750 4.120857 13.77065 0.192332 25.73495
750 3.906572 13.05457 0.182331 24.39674
800 3.512296 19.32817 0.564696 8.676591
800 4.261006 26.00972 1.017706 8.409003
800 4.434291 17.82804 1.617212 8.467262
850 7.028688 28.20471 3.924492 0.039359

145
Table C-4 1-Propylnaphthalene pyrolysis for 1 second

Temp
2-Methylnaphthalene 1-Methylnaphthalene 1-Ethylnaphthalene
(ºC)
550 0 0.04607 0.081546
550 0.023956 0.085227 0.127542
600 0 0.266626 0.620593
600 0.016849 0.429986 0.835062
650 0.083487 2.298122 3.446526
650 0.099037 3.382186 4.562416
650 0.098509 3.546215 5.03337
700 0.317276 10.96981 6.808998
700 0.306798 10.86871 6.787621
700 0.413572 13.69928 7.47555
750 0.568682 15.27528 4.325184
750 0.533195 14.50769 5.322541
750 0.524232 14.45906 5.583846

Table C-5 1-Propylnaphthalene pyrolysis for 1.9 second

Temp
2-Methylnaphthalene 1-Methylnaphthalene 1-Ethylnaphthalene
(ºC)

550 0 0.022859 0.15417


550 0 0.030889 0.076173
600 0.067467 0.714308 1.964784
600 0.066605 0.679285 1.716924
650 0.140119 4.198003 6.502596
650 0.159733 5.382501 7.542941
700 0.496235 14.21695 7.870947
700 0.356988 9.191363 4.439179
750 0.750024 17.76413 2.539105

146
12 APPENDIX D MATLAB CODE TO ESTIMATE KINETIC

PARAMETERS OF 2-METHYLNAPHTHALENE

FORMATION IN 1-METHYLNAPHTHALENE

PYROLYSIS

% The program is used to simulate the formation of 2-MN formation in 1-MN pyrolysis
% The data of temperature, 2-MN yield are known and shown in variable Temp and Product;
% Kinetic parameters of 2-MN are obtained and listed in the subroutine yPY_2

Temp = [800, 800, 800, 850, 850, 850, 900, 900, 900, 900, 925, 925, 950, 950, 950, 950, 950, 975, 1000, 1000]';
Temp = Temp +273;

Product = [0.786347702, 0.67761666, 0.878153382, 4.263037791, 3.438801569, 2.297939001, 10.94646119,


11.18857012, 11.26219899, 10.59127378, 11.71099403, 9.850634565, 5.763860861, 6.996588821, 7.163376581,
1.246933162, 0.288107225, 0.552685902]';

minnorm=1e6;
minfactor =[1,1,1];
minf1=minfactor;
numofIsOK=1;
options= statset('MaxIter', 200, 'TolFun',1e-8,'TolX',1e-8,'Display','off','FunValCheck','off');
% i is the initial estimate of A, pre-exponential parameter
for i=9:2:50
% j is for the initial estimate of Ea/R
for j=5000:2000:75000
% k is for Xf
for k=10:3:40
factor=[i, j, k]';
[factor1, r]=nlinfit(Temp, Product,'yPY_2',factor,options);
x=norm(r);

if ( minnorm>x )
minnorm=x;
minfactor=[i,j,k];
minf1=factor1;
end;
end
end
end
disp('Min norm of r is:');
disp(minnorm);
disp('Min factor combination is');
disp(minfactor);

147
disp('The corresponding factor is:');
disp(minf1);

% definition of subroutine yPY_2


% T is the temperature deg Kelvin
function yPY1=yPY_2(factor,T)
A1=factor(1);
Ea1=factor(2);
Xf=factor(3);

%2MN pyrolysis data


A2=26.64;
Ea2=31582;
A1E1 = exp(A1-Ea1./T);
A2E2 = exp(A2-Ea2./T);
yPY1 = Xf*A1E1.*(exp(-A1E1)-exp(-A2E2))./(A2E2-A1E1);

148

You might also like