136
MID-INFRARED LASER DIAGNOSTICS FOR CHEMICAL KINETICS STUDY OF OXYGENATES A DISSERTATION SUBMITTED TO THE DEPARTMENT OF MECHANICAL ENGINEERING AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Wei Ren August 2013

Ren, Wei (2013). Mid-infrared laser diagnostics for chemical kinetics

  • Upload
    donga

  • View
    215

  • Download
    0

Embed Size (px)

Citation preview

MID-INFRARED LASER DIAGNOSTICS FOR CHEMICAL

KINETICS STUDY OF OXYGENATES

A DISSERTATION

SUBMITTED TO THE DEPARTMENT OF MECHANICAL ENGINEERING

AND THE COMMITTEE ON GRADUATE STUDIES

OF STANFORD UNIVERSITY

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Wei Ren

August 2013

iv

v

Abstract

Biofuels are classified as renewable because the carbon present in the vegetable oil

or animal fat feedstocks originates from carbon dioxide already present in the atmosphere.

Biofuels also offer additional benefits such as reduced carbon monoxide, hydrocarbon

and particulate matter emissions and the potential to reduce the world’s intense

dependence on fossil fuels. One of the current focuses on biofuel-based energy systems is

the design of advanced energy conversion devices using complex reaction

mechanisms. The development of these mechanisms requires a large experimental

database to ensure accuracy of computational predictions.

Infrared laser-absorption diagnostics are widely used in combustion research for fast,

sensitive, and non-intrusive measurements of species concentration, temperature, and

pressure. The combination of shock-heating and species-specific laser absorption

provides a state-of-the-art test platform for studying chemical kinetics. This thesis

explores three new areas of laser diagnostic research: (a) mid-infrared diagnostics, (b)

sensing in multiphase flows, and (c) applications to shock tube chemical kinetics.

Carbon monoxide (CO) and carbon dioxide (CO2) are particularly significant

diagnostic targets for combustion systems, since they are the primary intermediate or

product in combustion, and their concentrations can be interpreted to indicate combustion

efficiency. Previous laser-based absorption sensors were mainly designed to exploit

commercial telecom diode lasers in the 1.3-1.6 m (near-infrared) wavelength region.

Recent developments in quantum-cascade (QC) laser technology, resulting in room-

temperature, high power (mW) and single-mode laser sources, allow access to much

stronger absorption bands of CO and CO2 in the mid-infrared region. The development of

a novel CO diagnostic near 4.7 m using QC laser was demonstrated as part of this thesis

work. Spectroscopic parameters of the selected transitions were determined via

vi

laboratory measurements in a shock tube over the 1100-2000 K range and also at room

temperature. The sensor was then tested in shock tube combustion measurements of

temperature and CO concentration time-histories to validate the sensor performance.

In many practical combustion systems, fuels are injected as liquid spray that quickly

evaporates at elevated temperatures. The interference caused by droplet scattering makes

the direct absorption measurements inaccurate. A tunable diode laser (TDL) sensor with a

detection bandwidth of 40 kHz was developed for measuring time-varying gas

temperature of CO2 during the evaporation of shock-heated hydrocarbon fuel aerosols.

Wavelength-modulation spectroscopy with 1f-normalized second-harmonic detection

(WMS-2f/1f) was used to probe R(28) and P(70) transitions in the v1+v3 combination

band of CO2 near 2.7 m. Application of this sensor for accurate temperature

measurement of evaporating n-dodecane aerosols was performed in an aerosol shock tube.

These recently developed mid-IR laser diagnostics were then applied in studying the

thermal decomposition of oxygenates by measuring species concentration time-histories

behind reflected shock waves. In the study of methanol pyrolysis, experimental

conditions covered temperatures of 1266 to 1707 K, pressures of 1.1 to 2.5 atm, and

initial fuel concentrations of 1% and 0.2% with argon as the bath gas. Pathway and

sensitivity analyses for methanol decomposition were performed, leading to rate constant

recommendations with improved model performance. In the study of methyl formate

pyrolysis, the reaction rate constants of the unimolecular elimination reaction (MF →

CH3OH + CO) were measured using a shock tube/laser diagnostic method over the range

of temperature 1261-1524 K, and pressure 0.3-5.2 atm. Methanol is the major

intermediate during MF pyrolysis, so incorporation of the modified rate constants in the

methanol sub-mechanism leads to improved predictions of the full methanol time-

histories at all temperatures. The kinetic implications of some aspects of the CO time-

histories and suggestions for further improving the predictive capabilities of these

mechanisms are discussed.

Finally, the thermal decomposition of three ethyl esters, ethyl formate (C3H6O2),

ethyl acetate (C4H8O2) and ethyl propanoate (C5H10O2), was studied behind reflected

shock waves using laser absorption of H2O, CO2 and CO. Experimental conditions

covered temperatures of 1301-1636 K, pressures of 1.48-1.72 atm, and reactant

vii

concentrations of 2000 ppm in argon. Recently developed mid-IR laser diagnostics for

H2O (2.5 m), CO2 (4.2 m) and CO (4.6 m) provide orders-of-magnitude greater

detectivity compared to previous near-IR absorption sensors. The experimental results

have highlighted the significant differences among these three ethyl esters: negligible

CO2 production during ethyl formate pyrolysis, very slow CO formation rate during ethyl

acetate pyrolysis, and nearly equal formation rate of all three species during ethyl

propanoate pyrolysis. Detailed kinetic modeling was performed to understand how the

difference in the alkyl length affects the fuel destruction pathways. Rate of production

and sensitivity analyses using the current kinetic models were also performed to interpret

the results. The experiments provide the first laser-based time-history measurements of

CO, CO2 and H2O during the pyrolysis of these potential bio-diesel surrogate fuels in a

shock tube.

viii

ix

Acknowledgements

Looking back at my PhD years at Stanford, there are many people I would like to

thank that have trusted me, helped me, and encouraged me. I owe many thanks to my

advisor Prof. Ronald Hanson for his guidance and support in this work. I enjoyed each

time meeting with him, presenting my research, discussing the results, and solving the

problems. What I learned from him and the world-class research will be remarkably

beneficial for my future career. I would also like to thank my reading committee,

Professor Tom Bowman and Dr. Dave Davidson, for suggestions regarding the content of

this thesis.

It is so fortunate for me to work with many outstanding people in the Hanson Group.

I am especially grateful to Dr. David Davidson for the contributions he has made to this

research and the arrangement of the experimental facilities making the lab a comfortable

place to work in. I am also grateful to Dr. Jay Jeffries for the technical contributions he

has made to this work. I would like to thank Professors Jennifer Wilcox and Reginald

Mitchell for participating in my oral exam committee.

I feel incredibly fortunate to have been surrounded by so many talented and friendly

people in my research group and at Stanford. I enjoyed the ski experience with labmates

to Lake Tahoe. Many thanks to alumni from the lab, Zekai Hong, Aamir Farooq, Xing

Chao and Jason Porter, for helping me start research when I initially joined the lab. I

would also like to thank my fellow students who made my life in lab more meaningful,

joyful, and certainly unforgettable: Brian Lam, Kai Sun, Ritobrata Sur, Mitchell Spearrin,

Shengkai Wang, Sijie Li, Yangye Zhu and coworkers. I am thankful to Haocheng

(Aerospace), Chunjing (Applied Physics), Yuan (Civil), Runzhi (Materials) and Kejie

(Physics), for their friendship and making my PhD life full of joys.

x

Most of all, I would like to thank my wife and family for their endless love and

support. Thank you to my mom, dad, mother-in-law and father-in-law for their never-

ending encouragement that helped me to reach for my dreams. I could not have

completed this work without my wonderful wife Erica, whom I met during this work and

whom I have created the small family. Our first son, Steven, was born in Stanford

Hospital and brought so much joy to our lives.

xi

Table of Contents

Abstract ........................................................................................................................v

Acknowledgements .......................................................................................................... ix

Table of Contents ............................................................................................................. xi

List of Tables ....................................................................................................................xv

List of Figures................................................................................................................ xvii

Chapter 1. Introduction ..................................................................................................1

1.1 Motivation and Background.....................................................................................1

1.2 Overview of Dissertation .........................................................................................4

Chapter 2. Mid-IR Laser Absorption Detection of Carbon Monoxide ......................7

2.1 Introduction ..............................................................................................................7

2.2 Fundamental Spectroscopy ......................................................................................8

2.3 Line Selection...........................................................................................................9

2.4 Spectroscopic Measurement and Verification .......................................................12

2.5 Sensor Validation in Shock Tube Experiments .....................................................18

2.5.1 Scanned-Wavelength CO Sensor Using a Single QC Laser ........................18

2.5.2 Fixed-Wavelength CO Sensor Using Two QC Lasers.................................21

2.6 Temperature and CO Concentration Measurements in Combustion Gases ...........23

Chapter 3. Two-Line Thermometry for Multiphase Combustion Flows .................27

3.1 Introduction ............................................................................................................27

3.2 Wavelength Modulation Spectroscopy Fundamentals...........................................28

3.3 Sensor Design.........................................................................................................32

3.3.1 Line Selection...............................................................................................32

3.3.2 Measurement Uncertainties..........................................................................32

xii

3.4 Temperature Measurement in CO2/Ar Gas ............................................................34

3.4.1 Experimental Setup ......................................................................................34

3.4.2 Experimental Results ...................................................................................35

3.4.3 Comparison of CO and CO2 Thermometry..................................................37

3.5 Sensor Validation in a Aerosol Flow Cell..............................................................38

3.6 Temperature Measurement in Shock-Heated Aerosol ...........................................41

Chapter 4. Thermal Decomposition of Methanol and Methyl Formate ...................45

4.1 Introduction ............................................................................................................45

4.2 Experimental ..........................................................................................................46

4.2.1 QC Laser Absorption of CO at 4.56 m ......................................................46

4.2.2 CO2 Laser Absorption of Methanol and Methyl Formate............................47

4.3 High-Temperature Methanol Pyrolysis ..................................................................47

4.4 High-Temperature Methyl Formate Pyrolysis .......................................................57

Chapter 5. Thermal Decomposition of C3-C5 Ethyl Esters.......................................67

5.1 Introduction ............................................................................................................67

5.2 Experimental ..........................................................................................................69

5.2.1 Shock Tube and Laser Diagnostics ..............................................................69

5.2.2 Experimental Results ...................................................................................70

5.3 Kinetic Modeling ...................................................................................................71

5.4 Discussion ..............................................................................................................74

5.4.1 Ethyl Formate Pyrolysis ...............................................................................74

5.4.2 Ethyl Acetate Pyrolysis ................................................................................79

5.4.3 Ethyl Propanoate Pyrolysis ..........................................................................84

Chapter 6. Summary and Future Directions...............................................................93

6.1 Summary of Results ...............................................................................................93

6.1.1 Mid-IR CO Sensor near 4.7 m...................................................................93

6.1.2 Two-Line Thermometry for Multiphase Flows ...........................................94

6.1.3 Methanol and Methyl Formate Decomposition Study .................................94

6.1.4 Ethyl Ester Decomposition Study ................................................................96

6.2 Recommendations for Future Work.......................................................................96

xiii

6.2.1 Shock Tube Measurements of Reaction Rate Constants ..............................96

6.2.2 Multi-Species Measurements in Large Oxygenates and Blends..................97

6.2.3 Kinetics of Oxygenated Fuel Thrust ............................................................99

Appendix A: Ethylene and Methanol Diagnostics using CO2 Gas Laser .................101

A.1 Ethylene Diagnostic at 10.532 m ......................................................................101

A.1.1 Experimental .............................................................................................101

A.1.2 High-Temperature Ethylene Absorption Cross-Section ...........................102

A.2 Methanol Diagnostic at 9.676 m.......................................................................105

A.2.1 Methanol Absorption Cross-Section .........................................................105

A.2.2 Two-Line Differential Absorption Measurement .....................................107

Reference ....................................................................................................................109

xiv

xv

List of Tables

Table 2.1 Candidate CO lines for the measurements of temperature and CO concentration

based on the HITRAN 2004 database [13]. ......................................................11

Table 2.2 Line-strength and broadening parameters for the CO transitions. Uncertainties

of measurements are given in the parentheses; the extrapolation of 2CO-Ar to

296 K following (Eqn. 2-4) is based on the experimental data over the

temperature range of 1100-2000 K. ..................................................................17

Table 3.1 Measured spectroscopic data for the selected CO2 line pair (from [40])...........31

Table 4.1 Summary of current methanol and methyl formate pyrolysis experiments. ......48

Table 4.2 Reaction rate constants (near 1 atma) used in the current study: k = ATnexp(-

Ea/RT) ................................................................................................................53

Table 4.3 Test conditions and rate constant data for reaction: CH3OCHO → CH3OH +

CO. ....................................................................................................................62

Table 5.1 Summary of reflected shock conditions for ethyl ester pyrolysis......................71

Table 5.2 EF pyrolysis submechanism; cm3/mol/sec/cal units. .........................................75

Table 5.3 EA pyrolysis submechanism; cm3/mol/sec/cal units. ........................................80

Table 5.4 Reaction rate constants modified in the Metcalfe et al. [83] mechanism;

cm3/mol/sec/cal units.........................................................................................87

Table 6.1 Stanford IR laser diagnostics for combustion gases ..........................................98

Table 6.2 New species and potential diagnostics in future ................................................98

Table A.1 Methanol absorption cross-section (m2/mol) at 1 atm and 297 K. .................106

xvi

xvii

List of Figures

Figure 1.1 World petroleum and other liquid fuel supply in three cases, 1990-2040

(million barrels per day); source: Annual Energy Outlook in 2013 [1].............2

Figure 1.2 Absorption line strengths of CO, H2O and CO2 at 1500 K (from HITRAN

database [13]).....................................................................................................3

Figure 2.1 Absorption line-strengths of CO at 1500K (from HITRAN 2004 database [13]).

............................................................................................................................8

Figure 2.2 Calculated spectra of 0.1% CO, 1% H2O and 1% CO2 in air under shock tube

combustion conditions: T = 1500 K, P = 1 atm, L = 10 cm.............................10

Figure 2.3 Temperature sensitivities (left-hand axis) and line-strength ratios (right-hand

axis) for two representative line pairs. Solid line: line pair A (v” = 1, R(21)

and v” = 0, R(12)) for single-laser scanned-wavelength temperature sensing;

dashed line: the v” = 1, R(21) and v” = 0, P(20) lines for dual-laser fixed-

wavelength temperature sensing (selected from the six individual lines listed

in Table 2.1). ....................................................................................................11

Figure 2.4 Calculated vibrational relaxation time (P = 1.5 atm) for CO-Ar, CO-He-Ar and

CO-H2-Ar mixtures (calculations from reference [30])...................................13

Figure 2.5 Experimental setup for the measurement of spectroscopic parameters of CO

transitions in a shock tube................................................................................13

Figure 2.6 Illustration of (a) the measured raw-data traces (pressure, transmission through

the shock tube and the etalon) of the R(12) transition at 2190.02 cm−1, taken at

2.5 kHz with 0.496% CO/1% H2/Ar mixtures behind the reflected shock

(vibrationally equilibrated reflected shock conditions: 1450 K, 1.63 atm); (b)

the reduced line-shape of the R(12) transition (solid line, top panel), its best-

fit Voigt profile (dashed line, top panel), and the residual (bottom panel). ....15

xviii

Figure 2.7 Comparison of the measured line-strengths for the CO transitions at high

temperatures with the HITRAN database [13]. ...............................................16

Figure 2.8 Ar-broadening coefficient 2CO-Ar measurements for the CO transitions: R(12),

R(13) and R(21). The two-parameter best fit extrapolated to 296 K gives 2CO-

Ar(296 K)=0.079±0.007 cm-1/atm and n = 0.581±0.012 for transition R(12),

2CO-Ar(296 K)=0.079±0.009 cm-1/atm and n = 0.600±0.016 for transition

R(13), and 2CO-Ar(296 K) = 0.072±0.007 cm-1/atm and n = 0.571±0.012 for

transition R(21), respectively...........................................................................16

Figure 2.9 Room-temperature (296 K) spectroscopic parameter measurements for (a)

line-strength using the measured integrated absorbance versus P1 (20-60 Torr),

and (b) Ar-broadening coefficient using the measured collisional FWHM

versus P1...........................................................................................................18

Figure 2.10 Simulated peak absorbance ratio for the line pair R(21)/R(12) and R(21)/P(20)

using the spectroscopic parameters listed in Table 2.2....................................19

Figure 2.11 Sample traces of laser transmission and pressure (top panel), as well as

absorbance and temperature (bottom panel) measured in non-reactive test

gases (0.49% CO/ 2% H2 /Ar, vibrationally equilibrated reflected shock

conditions: 1526 K, 1.57 atm). A single QCL was used to scan over the line

pair R(21) and R(12) at 2.5 kHz. .....................................................................20

Figure 2.12 Shock tube validation measurements for the scanned-wavelength (measured

for a single scan behind the reflected shock, solid squares) and the dual-laser

fixed-wavelength (averaged over the first 0.3-1 ms after the shock, solid

triangles) direct absorption CO sensors (0.49% CO/2% H2/Ar, 1.3-1.7 atm).20

Figure 2.13 Experimental setup for the fixed-wavelength two-line temperature and CO

concentration measurements in a shock tube...................................................21

Figure 2.14 Fixed-wavelength temperature measurements using two QC lasers with 0.49%

CO/2% H2/Ar: (a) measured absorbance traces for the two lasers; (b)

measured temperature and pressure. Vibrationally equilibrated reflected shock

conditions: P5 = 1454 K, T5 = 1.62 atm. ..........................................................22

xix

Figure 2.15 Absorbance time-histories of R(21) and P(20) during the pyrolysis of methyl

formate. Initial reflected shock conditions: T5 = 1364 K, P5 = 1.63 atm, 0.5%

MF/Ar. .............................................................................................................23

Figure 2.16 Temperature and CO concentration measured during a shock with initial

mixture of 0.5% MF/Ar; simulations using the Dooley et al. [33] mechanism

are shown for comparison. Initial reflected shock conditions: T5 = 1364 K, P5

= 1.63 atm. .......................................................................................................24

Figure 2.17 Temperature and CO concentration measurements during MF oxidation for a

mixture of 0.494% MF, 0.988% O2 (= 1) and Ar; simulations using the

Dooley et al. [33] mechanism are shown for comparison. Initial temperature

and pressure behind the reflected shock are T5 = 1379 K, P5 = 1.67 atm........24

Figure 3.1 Calculated CO2 (1%) absorption spectra for R(28) transition at 3633.08 cm-1

(2752.48 nm) and P(70) transition at 3645.56 cm-1 (2743.06 nm) under typical

shock tube conditions: T2 = 650 K, P2 = 0.5 atm; T5 = 1200 K, P5 = 1.0 atm; L

= 10 cm. ...........................................................................................................31

Figure 3.2 Calculated extinction cross section (Mie scattering code [47]) for liquid n-

dodecane droplets; Dm is the median diameter of aerosol droplet size. ...........31

Figure 3.3 WMS simulation for (a) 2f/1f magnitude of 2752nm line and (b) 2f/1f ratio of

2752nm/2743nm line pair, as a function of temperature for specified pressures

at optimized modulation depths; 2% CO2 in Ar, T = 900-1600 K, L = 10 cm.

..........................................................................................................................33

Figure 3.4 WMS simulation for (a) 2f/1f magnitude of 2752nm line and (b) 2f/1f ratio of

2752/2743 line pair, as a function of temperature for specified CO2

concentrations; P = 0.5 atm, T = 400-900 K, L = 10 cm. ................................33

Figure 3.5 Shock tube experimental setup. ........................................................................34

Figure 3.6 Measured temperature and pressure trace during a shock with CO2/Ar mixture

without aerosol. Initial conditions: P1 = 55.0 Torr, T1 = 298 K; incident shock

(calculated): P2 = 0.43 atm, T2 = 697 K; reflected shock (calculated): P5 =

1.48 atm, T5 = 1199 K......................................................................................36

xx

Figure 3.7 Temperatures measured by the WMS-2f/1f sensor in shock-heated CO2/Ar

mixture without aerosol versus calculated values using shock jump equations;

±1.5% error bars. The square points represent T5 behind reflected shocks (P5 =

1.0-1.5 atm); the triangular points represent T2 behind incident shocks (P2 =

0.4-0.6 atm). .....................................................................................................36

Figure 3.8 Measured temperature and pressure trace at 70 cm from the endwall with

CO2/Ar mixture. Initial conditions: P1 = 50.1 Torr, T1 = 298 K; incident shock

(calculated): P2 = 0.35 atm, T2 = 649 K...........................................................37

Figure 3.9 Aerosol flow cell experimental setup. ..............................................................38

Figure 3.10 Measured WMS- (a) 2f, (b) 1f and (c) 2f/1f signals in an aerosol flow cell for

the R(28) transition of CO2 with different aerosol loadings; v represents the

droplet extinction coefficient. ..........................................................................40

Figure 3.11 Comparison of the measured WMS-2f/1f data with the simulated value under

the condition of no droplet scattering. .............................................................40

Figure 3.12 Measured temperature for an incident shock-heated aerosol with the WMS-

2f/1f sensor located at 10 cm from the endwall: (P2)w/o evap = 0.50 atm, (T2)w/o

evap = 558 K; (P2)post evap = 0.54 atm, (T2)post evap = 528 K; P5 = 1.79 atm, T5 =

796 K. A non-resonant 660 nm laser is used to indicate the droplet scattering.

..........................................................................................................................41

Figure 3.13 Temperatures measured in aerosol shock tube by the WMS-2f/1f sensor

versus calculated values using numerical code; ±1.8% error bars. The square

points represent T5 behind reflected shocks (P5 = 1.0-1.5 atm); the triangular

points represent the post-evaporation T2 behind incident shocks (P2 = 0.4-0.6

atm). .................................................................................................................42

Figure 4.1 Measured (solid lines) and simulated (dashed lines) methanol and CO

concentration time-histories during the pyrolysis of methanol (time-zero:

arrival of the reflected shock wave). Simulations used the Li et al. [17]

mechanism. The initial post-shock temperature and pressure are indicated....48

Figure 4.2 Comparison of (a) methanol and (b) CO concentration time-histories with

different absorption cross-sections in Beer’s law. ...........................................50

xxi

Figure 4.3 Comparison of the measured (a) methanol and (b) CO time-histories with a

detailed chemical kinetic model. Long-dashed lines: predictions of the Li et al.

[17] model at the nominal temperature shown; short-dashed lines: computed

uncertainty bounds due to ±15 K uncertainty in the T5 value..........................50

Figure 4.4 Sensitivity analysis (unmodified Li et al. [17] mechanism) of CH3OH at 100

s for 1% methanol in argon at 1458 and 1567 K, respectively......................51

Figure 4.5 Influence of modified k1 (branching ratio from Jasper et al. [53]) on the (a)

CH3OH and (b) CO predictions by the Li et al. [17] mechanism. The spike at t

= 0 is a result of beam steering from the detector during the passage of the

reflected shock and is not kinetic in nature......................................................52

Figure 4.6 Flux analysis of methanol pyrolysis at 800 s for 0.2% CH3OH/Ar at 1623 K

and 1.1 atm.......................................................................................................54

Figure 4.7 Reaction rate constants of CH3OH+H (k2) and branching ratio. ......................55

Figure 4.8 Effect of modifications to the Li et al. [17] mechanism predictions for (a)

CH3OH and (b) CO concentration time-histories during the pyrolysis of

methanol...........................................................................................................56

Figure 4.9 Sensitivity analysis (Li et al. [17] mechanism with k1 modified) of CO

concentration at 100 s for 0.2% CH3OH/Ar at 1507 K and 1623 K..............57

Figure 4.10 Representative CO concentration time-histories measured during the

decomposition of MF at various temperatures under a fixed initial fuel

concentration (0.1% MF/Ar) compared with the predictions of the Dooley et

al. [33] mechanism and that with k3a-k3c revised from Metcalfe et al. [73]. ....58

Figure 4.11 Local sensitivity analysis for CO concentration using the Dooley et al. [33]

mechanism (0.1% MF/Ar, 1376 K, 1.58 atm). ................................................59

Figure 4.12 Example MF decomposition k3a rate constant determination. Solid black line,

experimental data; solid red line, best fit to the data with the optimal value of

k3a; dashed lines, variation of k3a±50%. ...........................................................60

Figure 4.13 Comparison of measured k3a (1.5-1.7 atm) with previous rate constants

(LLNL [75], Princeton [33], Argonne [74] and NUI [73]) for the reaction

xxii

CH3OCHO → CH3OH + CO. Least-squares fit (in black) to experimental

data gives k3a = 1.11013 exp(-29556/T, K) s-1.................................................60

Figure 4.14 Summary of recent studies of k3a. Symbol: shock tube measurement; dashed

line: Peukert et al. [74]; dash-dot line: Metcalfe et al. [73]. ............................61

Figure 4.15 Comparisons of measured and simulated methanol time-histories for 1%

methyl formate in argon. Only the reaction rate constants k1, k2 and k3a are

modified in the Dooley et al. [33] mechanism.................................................63

Figure 4.16 Example CO concentration time-histories: solid line, measurement; dashed

line, simulation using unmodified Dooley et al. [33] mechanism; dash-dot

line, simulation using the Dooley et al. mechanism with k1, k2 and k3a modified.

..........................................................................................................................63

Figure 4.17 CO sensitivity (Dooley et al. [33] mechanism) for 0.1% MF/Ar, 1607 K and

1.5 atm..............................................................................................................64

Figure 4.18 Reaction rate constants k3a, k3b and k3d in the Dooley et al. [33] mechanism.64

Figure 4.19 Effect of modifications to the Dooley et al. [33] model predictions for the CO

concentration time-histories during the pyrolysis of methyl formate. .............66

Figure 4.20 Comparisons of measured and simulated methanol time-histories during MF

pyrolysis. ..........................................................................................................66

Figure 5.1 The molecular structures of (a) ethyl formate (b) ethyl acetate and (c) ethyl

propanoate........................................................................................................68

Figure 5.2 Measured species time-histories during the pyrolysis of (a) EF (b) EA and (c)

EP at temperature near 1450 K and pressures near 1.5 atm, with fuel

concentration 2000 ppm in argon. ...................................................................70

Figure 5.3 Measured product fractional yield for (a) EF (b) EA and (c) EP at t = 1 ms. ..71

Figure 5.4 EF pyrolysis: major destruction pathways. ......................................................73

Figure 5.5 EA pyrolysis: major destruction pathways.......................................................74

Figure 5.6 EP pyrolysis: major destruction pathways. ......................................................74

Figure 5.7 Measured H2O and CO concentration time-histories during the pyrolysis of

ethyl formate. ...................................................................................................75

xxiii

Figure 5.8 Comparison of the measured (a) CO, (b) H2O and (c) CO2 concentration time-

histories with the model predictions during the pyrolysis of 2000 ppm EF in

argon: solid line, measurement; dashed line, simulation in this study.............77

Figure 5.9 ROP and sensitivity analyses of CO: 2000 ppm EF/Ar, 1500 K and 1.5 atm..77

Figure 5.10 H2O sensitivity: 2000 ppm EF/Ar, 1500 K and 1.5 atm.................................78

Figure 5.11 CO2 sensitivity: 2000 ppm EF/Ar, 1500 K and 1.5 atm.................................78

Figure 5.12 Comparison of the measured CO concentration time-histories during the

pyrolysis of EF, EA and EP; pressure near 1.5 atm, fuel concentration 2000

ppm. .................................................................................................................79

Figure 5.13 Comparison of the measured (a) CO, (b) H2O and (c) CO2 concentration

time-histories with the model predictions for 2000 ppm EA/Ar: solid line,

measurement; dashed line, simulation. ............................................................81

Figure 5.14 (a) ROP and (b) sensitivity analyses (using the current EA mechanism) of CO

during the pyrolysis of 2000 ppm EA/Ar at 1500 K and 1.5 atm. ...................83

Figure 5.15 CO2 sensitivity during the pyrolysis of 2000 ppm EA/Ar at 1500 K and 1.5

atm....................................................................................................................84

Figure 5.16 H2O sensitivity during the pyrolysis of 2000 ppm EA/Ar at 1500 K and 1.5

atm....................................................................................................................84

Figure 5.17 Measured (symbol-solid line) and simulated (dashed line, Metcalfe et al. [83])

CO, H2O and CO2 yields for 2000 ppm EP/Ar mixture at 1 ms. Temperature:

1301-1580 K; pressure: 1.4-1.7 atm. ...............................................................85

Figure 5.18 Main reaction pathways for EP pyrolysis using the Metcalfe et al. [83]

mechanism: 2000 ppm EP/Ar, 1350 K, 1.5 atm, at t = 200s. ........................85

Figure 5.19 Comparison of measured (a) H2O and (b) CO2 and (c) CO concentration

time-histories with the model predictions during the pyrolysis of 2000 ppm

EP/Ar. Solid line, current measurement; dash-dot line, simulation using the

Metcalfe et al. [83] mechanism; dashed line, simulation using the modified

Metcalfe et al. mechanism. ..............................................................................88

Figure 5.20 CO (a) ROP and (b) sensitivity analyses using the modified Metcalfe et al.

[83] mechanism: 2000 ppm EP/Ar, 1500 K and 1.5 atm.................................90

xxiv

Figure 5.21 CO2 sensitivity analysis using the modified Metcalfe et al. [83] mechanism:

2000ppm EP/Ar, 1500 K and 1.5 atm..............................................................90

Figure A.1 Schematic of CO2 laser diagnostic in shock tube measurements; ND: neutral

density filter, NBP: narrow bandpass filter....................................................102

Figure A.2 Pressure and laser absorbance time-histories for a nonreactive mixture: 1%

C2H4/Ar. Schlieren spikes caused by the density gradient across the shock

waves..............................................................................................................104

Figure A.3 Ethylene cross-sections (10.532m): 643-1959 K and 0.3-18.6 atm. Upper

panel: measured absorption cross section,meas; lower panel: comparisons of

meas with fit calculated using (Eqn. A-1).....................................................104

Figure A.4 Ethylene cross section (1.8-5.5 atm) as a function of temperature; best fit

using (Eqn. A-2).............................................................................................104

Figure A.5 IR spectra of gas-phase methanol at 298 K (from PNNL [115])...................105

Figure A.6 Determination of methanol absorption cross-section (297 K) at 9.676 m: (a)

measured methanol absorbance as a function of pressure (10-70 Torr); (b)

measured methanol cross-section as a function of pressure (data for

comparison at 1 atm are from Molina et al. [116], Sharpe et al. [115] and

Loper et al. [58]). ...........................................................................................106

Figure A.7 Methanol cross-sections (m) measured over 665-1940 K and 0.4-2.7

atm. Curve fit is given by (Eqn. A-4). ...........................................................107

Figure A.8 Laser absorbance data for 1% MF/Ar mixture at 1327 K and 1.5 atm..........108

Figure A.9 Measured absorption cross-sections of (a) methanol and (b) methyl formate at

wavelengths of 9.676 and 9.229 m; P = 0.6-2.7 atm...................................108

1

Chapter 1. Introduction

1.1 Motivation and Background

Energy demand around the world is continuously increasing, including petroleum,

the major source of fuel used in the transportation sector. According to the 2013 Annual

Energy Outlook from the U.S. Energy Information Administration (EIA), the world use

of petroleum and other liquid fuels will increase from 89.1 million barrels per day in

2012 to 111.8 million barrels per day in 2040, as shown in Figure 1.1 [1]. Due to the

progressive depletion of oil reserves and the negative environmental impact of fossil fuel

use, there are strong reasons for the development of advanced combustion systems with

higher efficiency and lower emissions, as well as the development of alternative sources

of energy. On the one hand, optimization of the current engine technologies can be

facilitated with accurate predictive models describing the combustion phenomena that

occur within the engine. On the other hand, biofuels, especially bioalcohol and biodiesel,

are among the most viable liquid transportation fuels for the foreseeable future and can

contribute significantly to sustainable development in terms of economic and

environmental concerns. The combustion characteristics of these new types of fuels need

to be fully understood before their usage in engine systems.

The complexity of practical fuels makes it impossible to include all of their

components, either in an experimental test or in a computational model. Therefore, fuel

surrogates are often employed as alternatives to study the chemistry of the fuels of

interest. In this thesis, research work has focused on using state-of-the-art laser

absorption diagnostics and shock tube methods to investigate the chemical kinetics of

2

small oxygenate compounds (methanol, methyl/ethyl ester), which are treated as biofuel

surrogates.

Figure 1.1 World petroleum and other liquid fuel supply in three cases, 1990-2040 (million barrels per day);

source: Annual Energy Outlook in 2013 [1].

Laser absorption spectroscopy techniques play a large and growing role in the

measurement of flow-field parameters such as temperature, gas composition, velocity,

and pressure [2–5]. These sensors are highly attractive for combustion and propulsion

applications due to their non-intrusive, in situ and line-of-sight measurements with fast

time response. Most of the combustion products have absorption spectra in the infrared

(IR) region as illustrated in Figure 1.2, where the absorption line strengths of CO, H2O,

and CO2 are plotted as a function of wavelength from 1 to 6 m at a representative

combustion temperature of 1500 K. Previous IR-laser-based absorption sensors were

mainly designed to exploit commercial telecom diode lasers in the 1.3-1.6 m (near-IR)

wavelength region [6–12]. These overtone and combination absorption bands are orders-

of-magnitude weaker compared to the fundamental bands between 2.5 and 6 m (within

mid-IR region) as shown in Figure 1.2. While the near-IR sensors can take advantage of

current optical fiber technology for applications such as wavelength-multiplexing, the

relatively small absorption strength of transitions in this region limit the application of

these sensors to high concentration or long path-length problems.

60.00

80.00

100.00

120.00

1990 2000 2010 2020 2030 2040

Low Price Oil

High Price Oil

2011History Projections

Reference

3

Figure 1.2 Absorption line strengths of CO, H2O and CO2 at 1500 K (from HITRAN database [13]).

In recent years, developments in quantum-cascade (QC) laser technology, resulting

in room-temperature, high power (mW) and single-mode laser sources, have enabled

access to these stronger absorption bands of combustion gases in the mid-IR region [14–

16]. Absorption sensors at these longer wavelengths offer greater sensitivity and potential

for more accurate and precise measurements than was possible previously. As part of this

thesis work, we developed the first mid-IR CO and CO2 diagnostics (near 4.6 and 4.3 m)

for high-temperature combustion applications. Combined with a previously developed

H2O sensor (near 2.5 m), these mid-IR laser diagnostics have been applied in studying

shock tube chemical kinetics in this thesis.

High-temperature chemical kinetics experiments, such as measurements of ignition

delay times, rate constants of elementary reactions and species concentration time-

histories, are regularly performed behind reflected shock waves in ultra-clean shock tubes.

Shock tubes are nearly ideal devices for studying chemical kinetics as they provide well-

controlled step changes in temperature and pressure. For moderate or large diameter

tubes, the uniform conditions behind the reflected shock waves are generally not

significantly affected by surface or transport phenomena. The combination of shock-

heating and species-specific laser detection provides a state-of-the-art test platform for

studying chemical kinetics. The experimental results enable one to follow the time

sequence of events occurring in a highly-complex reaction process: from initial fuel

4

breakdown, intermediates/radical build up during and after the induction period, finally to

the formation of the combustion products.

The kinetic target of this thesis is to understand the pyrolysis of oxygenate

compounds, including methanol, methyl formate and C3-C5 ethyl esters (ethyl formate,

ethyl acetate, ethyl propanoate). These oxygenates are either treated as biofuel surrogates

or found to be crucial intermediates during the combustion of other important

hydrocarbon fuels. As discussed before, pyrolysis is the initial step of combustion and

thus the pyrolytic behavior must be well-characterized to accurately describe the

oxidation of a fuel. In this thesis, multi-species concentration time-histories were

measured during the pyrolysis of these oxygenate compounds to provide the rate constant

determination of several important elementary reactions and the validation of the detailed

chemical kinetic mechanisms. Such experimental data obtained using the shock tube/laser

diagnostics will undoubtedly be valuable to guiding future model development.

1.2 Overview of Dissertation

The dissertation is aiming to describe and discuss the key advancements achieved in

the relevant work, which are divided into the next four chapters accordingly:

1) Chapter 2 presents a mid-IR absorption sensor developed for measuring carbon

monoxide and temperature using CO transitions in the fundamental vibrational band near

4.7 m. It includes the introduction of the fundamental theory of laser absorption

spectroscopy and the direct absorption (DA) diagnostic method. Selection of optimal

transitions, measurements of spectroscopic parameters, and validation of the sensor in a

shock tube are discussed in this chapter.

2) Chapter 3 presents the development of a tunable diode laser (TDL) sensor near

2.7 m for measuring gas temperature of CO2 in shock-heated evaporating aerosols. In

many practical combustion processes, fuels are injected as liquid spray which quickly

evaporates at elevated temperatures. A normalized wavelength modulation spectroscopy

with second-harmonic detection (WMS-2f/1f) method is demonstrated to eliminate the

5

interference from droplet scattering. Applications of this sensor for accurate temperature

measurement of evaporating n-dodecane aerosols are performed in an aerosol shock tube.

3) Chapter 4 describes the thermal decomposition study of methanol and methyl

formate by measuring methanol and CO concentration time-histories behind reflected

shock waves. Pathway and sensitivity analyses for methanol decomposition were

performed, leading to rate constant recommendations for improved model performance of

the Li et al. [17] mechanism. In the study of methyl formate (MF) pyrolysis, the reaction

rate constants of the unimolecular elimination reaction (MF → CH3OH + CO) are

measured using the shock tube/laser diagnostic method.

4) Chapter 5 describes the thermal decomposition of three ethyl esters, ethyl formate

(C3H6O2), ethyl acetate (C4H8O2) and ethyl propanoate (C5H10O2) by measuring H2O,

CO2 and CO concentration time-histories behind reflected shock waves. Recently

developed mid-IR laser diagnostics for H2O (2.5 m), CO2 (4.2 m) and CO (4.6 m)

provide orders-of-magnitude greater detectivity compared to previous near-IR absorption

sensors. Detailed kinetic modeling is performed to understand how the difference in the

alkyl length affects the fuel destruction pathways. Rate of production and sensitivity

analyses were also performed to interpret the results.

Finally, Chapter 6 summarizes the major advancements of the work in this thesis

and suggests future research directions.

6

7

Chapter 2. Mid-IR Laser Absorption

Detection of Carbon Monoxide

2.1 Introduction

Laser absorption spectroscopy techniques play a large and growing role in the

measurement of flow-field parameters such as temperature, gas composition, velocity,

and pressure [2–5]. These sensors are highly attractive for combustion and propulsion

applications due to their non-intrusive nature, fast time response, and in situ measurement

capability. Carbon monoxide (CO) is a particularly significant target for hydrocarbon-

fueled systems, since it is a toxic pollutant from combustion devices and a primary

product of incomplete combustion, and its concentration can be interpreted to indicate

combustion efficiency.

The absorption spectra of CO, H2O and CO2 in the near- to mid-infrared region at

1500 K are illustrated in Figure 2.1, where the absorption line-strengths (from the

HITRAN 2004 database [13]) are plotted as a function of wavelength from 1-6 m. The

fundamental band of CO holds the most promising candidate transitions owing to their

much stronger line-strengths and relatively weaker interference from other combustion

species. Work has been reported using transitions in three different vibrational bands of

CO: the second overtone band (v = 3) near 1.55 m [6,8,18], the first overtone band (v

= 2) near 2.3 m [19–22], and the fundamental band (v = 1) near 4.6 m [23–28]. The

absorption strength of the fundamental band is approximately 104 and 102 times stronger

compared to the overtone bands near 1.55 m and 2.3 m, respectively, making it

promising for sensitive detection with relatively low CO concentration and/or short path

length.

8

Figure 2.1 Absorption line-strengths of CO at 1500K (from HITRAN 2004 database [13]).

Developments in quantum-cascade (QC) laser technology, resulting in room-

temperature, relatively high power (mW), narrow line-width, and single-mode QC lasers,

have led to broad applications of these sources in high-resolution spectroscopy and high-

sensitivity detection of trace gases [14–16]. In this thesis, we discuss the development of

cw DFB-QCL-based mid-IR absorption of CO for in situ detection in combustion gases

and specifically in a shock tube. Sensors for temperature and CO concentration

measurements using scanned-wavelength direct absorption (DA) with a single room-

temperature QC laser and using fixed-wavelength DA with dual QC lasers are both

developed to provide fast and flexible diagnostics for different applications.

2.2 Fundamental Spectroscopy

The fundamental theory governing the light transmission through gaseous species is

the Beer-Lambert law. When spectrally narrow radiation at frequency v passes through a

uniform gas medium of length L [cm], the transmitted intensity It is related to the incident

intensity I0 by the Beer-Lambert law:

0

exp( ),ti v

v

ISPx L

I

Eqn. 2-1

where S [cm-2atm-1] is the line-strength of the specific transition, P [atm] is the total

pressure, xi is the mole fraction of the absorbing species i, and v [cm] is the line-shape

function. The dimensionless product v = SPxivL is defined as absorbance, with kv =

9

SPxiv the absorption coefficient. Since the line-shape function v is normalized to have

unit area across the line, the integrated absorbance can be expressed as

( ) .i v i iA dv S T Px L Eqn. 2-2

The Voigt line-shape function v combines both temperature and collisional

broadening. The collision-broadened full-width at half maximum (FWHM) of the

absorbing species i is

-1, cm 2 ,c j jij

v P x Eqn. 2-3

where xj is the mole fraction of the collisional partner j, and 2ji [cm-1atm-1] is the

broadening coefficient of j with i. From an experimental point of view, it is of practical

interest to have a simple model of the variation of the FWHM with temperature, typified

by the following commonly-used expression:

002 ( ) 2 ,

nT

T TT

Eqn. 2-4

where T0 is the reference temperature (usually 296 K) and n is the temperature coefficient.

The line-strength S [cm-2atm-1] has a temperature dependence:

1

0 0 0 00

0 0

( ) " 1 1( ) ( ) exp 1 exp 1 exp ,

( )

Q T T hcv hcvhcES T S T

Q T T k T T kT kT

Eqn. 2-5

where Q(T) is the partition function, E” [cm-1] is the lower-state energy, v0 [cm-1] is the

line-center frequency, and h, c, k are Planck's constant, speed of light and Boltzmann’s

constant, respectively. The absorption measurement of temperature is commonly based

on a two-line technique [24]. Temperature is inferred from the ratio of the integrated

absorbance under the absorption feature or the line-center absorbance of two molecular

transitions of the same species.

2.3 Line Selection

Absorption spectra of the CO fundamental band between 4.3 and 5.8 m were

computed based on the HITRAN database [13] for typical shock tube combustion

10

conditions (1000-2000 K, 1 atm, 0.1% CO/1% H2O/1% CO2) to find suitable CO

transitions. A systematic line-selection procedure was used to find lines with sufficient

absorption strength, isolation from interfering absorption, temperature sensitivity, and the

availability of the commercial laser sources [29].

Two cw, room-temperature, DFB-QC lasers were subsequently acquired from Alpes

Lasers SA to access the R-branch near 4.6 m and the P-branch near 4.8 m of the

fundamental band of CO, respectively. For the laser frequency ranges of 2048.6 to 2061.3

cm-1 and 2185.8 to 2200.3 cm-1, three sets of closely spaced line pairs were selected for

single-laser, scanned-wavelength temperature sensing: line pair A (v” = 0, R(12) and v” =

1, R(21) near 2190 cm-1), line pair B (v” = 0, R(13) and v” = 1, R(22) near 2194 cm-1),

and line pair C (v” = 0, P(20) and v” = 1, P(14) near 2060 cm-1). Their spectroscopic

parameters (for line pairs A, B and C) from the HITRAN database [13] are summarized

in Table 2.1. A spectral simulation of 0.1% CO in air (T = 1500 K, P = 1 atm, L = 10 cm)

for these three line pairs is illustrated in Figure 2.2, along with the interfering absorption

of 1% H2O and CO2. It should be noted that the interference from H2O and CO2 is mostly

negligible at these wavelengths under the shock tube conditions of interest.

Figure 2.2 Calculated spectra of 0.1% CO, 1% H2O and 1% CO2 in air under shock tube combustion

conditions: T = 1500 K, P = 1 atm, L = 10 cm.

11

Table 2.1 Candidate CO lines for the measurements of temperature and CO concentration based on the

HITRAN 2004 database [13].

Figure 2.3 Temperature sensitivities (left-hand axis) and line-strength ratios (right-hand axis) for two

representative line pairs. Solid line: line pair A (v” = 1, R(21) and v” = 0, R(12)) for single-laser scanned-

wavelength temperature sensing; dashed line: the v” = 1, R(21) and v” = 0, P(20) lines for dual-laser fixed-

wavelength temperature sensing (selected from the six individual lines listed in Table 2.1).

These three line pairs, each with CO transitions from two different vibrational levels,

have a spectral separation of 0.4-1.5 cm-1, within the typical 2 cm-1 rapid-tuning range of

the commercial QC lasers. Representative analyses of the line-strength ratio and

temperature sensitivity for the line pair A are plotted as solid lines in Figure 2.3.

Typically the line-strength ratio should not be too far from unity. The sensitivity, defined

here as the unit change in line-strength ratio for a unit change in temperature should at

least be 1 for sensitive temperature measurements. These two curves suggest this line pair

can be used for accurate temperature sensing at elevated temperatures, e.g., between 1000

and 3000 K.

Line pair Transition

(v”, J”)

Frequency

(cm-1)

Wavelength

(nm)

Separation

(cm-1)

S @ 296 K

(cm-2atm-1)

E”(cm-1)

A0, R(12) 2190.02 4566.17

1.487.13 299.77

1, R(21) 2191.50 4563.08 4.32×10-5 3022.09

B0, R(13) 2193.36 4559.22

1.106.04 349.70

1, R(22) 2194.46 4556.93 3.02×10-5 3105.65

C0, P(20) 2059.91 4854.58

0.428.76×10-1 806.38

1, P(14) 2060.33 4853.59 2.64×10-4 2543.06

12

Two-line thermometry, achieved by scanning two neighboring transitions with a

single laser, enables a relatively simpler system with lower cost. However, the tuning rate

of the QC lasers limited the sensor bandwidth to several kHz. High-temperature chemical

kinetic studies in a shock tube, where chemical reactions happen within milliseconds,

require a faster sensor, with 100 kHz bandwidth or greater. Thus, a dual-laser, fixed-

wavelength method was pursued to provide highly time-resolved measurements. We

selected the v” = 0, P(20) and v” = 1, R(21) lines from the six individual lines listed in

Table 2.1 as the optimum line pair for temperature measurement using two different QC

lasers. The corresponding line-strength ratio and temperature sensitivity for this line pair

are shown as dashed lines in Figure 2.3.

2.4 Spectroscopic Measurement and Verification

The fundamental spectroscopic parameters such as line-strength, self- and air-

broadening coefficients of CO can be found in the HITRAN database [13]. However,

argon instead of air is usually used as the bath gas in shock tube kinetic studies.

Accordingly, there is need to investigate the Ar-broadening coefficient of each line and

its temperature dependence. Moreover, the validation of CO line-strength at high

temperature is essential for the accurate measurements as the measured absorbance is

compared with the simulation to infer gas mole fraction and temperature.

All spectroscopic measurements were performed in a 15.2 cm diameter stainless-

steel high-purity shock tube. The incident shock wave propagates through the tube,

raising the temperature and pressure of the test gas from (T1, P1) to (T2, P2). When the

shock wave reaches the end-wall of the tube, it is reflected and further elevates the

temperature and pressure of the test gas to (T5, P5). The gas temperature and pressure

immediately behind the shock wave can be calculated accurately using standard normal-

shock relations and the measured incident shock speed, with an uncertainty of ~1% in

temperature over the high-quality test time of 2 ms. Research grade gases (argon, helium,

and hydrogen >99.999%; 0.5% CO/Ar mixture with uncertainty <0.1%) were supplied by

Praxair Inc. Due to the significant time for CO to vibrationally relax behind the reflected

shock wave, a small portion of H2 (1%) is added to the 0.5% CO/Ar mixture to accelerate

13

the vibrational relaxation; see Figure 2.4 for the evaluation. The test mixtures were

manometrically prepared in a turbo-pumped stainless-steel mixing tank (40 L) with a

magnetically driven stirrer.

Figure 2.4 Calculated vibrational relaxation time (P = 1.5 atm) for CO-Ar, CO-He-Ar and CO-H2-Ar

mixtures (calculations from reference [30]).

Figure 2.5 Experimental setup for the measurement of spectroscopic parameters of CO transitions in a

shock tube.

A schematic of the experimental setup is demonstrated in Figure 2.5. The room-

temperature operated QC laser (Alpes Lasers) used for these measurements was

thermoelectrically cooled and housed with collimation optics in a sealed laser housing

Shock-heated gases

DiagnosticBeam

Iris

NBP

Wav

emet

er

I0 I

QCL

LaserController

FunctionGenerator

Solid etalon

Flip mirror

Beamsplitter

14

(Alpes HHL-L module). In addition, a laboratory water-cooled heat sink was installed to

the laser housing to achieve more stable laser performance. The laser wavelength is tuned

by varying the injection current and base temperature, which are controlled by a

combination of commercial temperature and injection current controllers (Alpes Lasers

TCU 200 and ILX Lightwave LDX-3232). The laser wavelength is rapidly tuned (1-10

kHz scan rate) over the desired absorption feature with a linear ramp of current from a

function generator. A ZnSe beam splitter was used to split the collimated laser beam (20-

40 mW) into two arms to be received by a pair of matched TE-cooled IR photovoltaic

detectors (Vigo Systems, 1 MHz bandwidth); one beam passes through the test gas of

15.2 cm path length in the shock tube, while the other propagates through a 7.6 cm long

solid germanium etalon in the ambient air. The etalon with a free spectral range (FSR) of

0.016 cm-1 enables the conversion of scan time to relative wavelength. A narrow-

bandpass IR filter (half power bandwidth 50 nm) was used to filter out emission and

unwanted ambient light. Before each shock tube experiment, the laser wavelength was

tuned to the desired transition by monitoring the absolute wavelength using a free-space

mid-IR wavemeter (Bristol 621).

The laser wavelength was typically tuned over a range of ~1 cm-1 at a frequency of

2.5 kHz, while the detector signal was sampled at 10 MHz to fully capture the absorption

feature. The data acquisition system was triggered by the pressure transducer located at 2

cm from the shock tube end-wall to record pressure and transmission signal (It) of the

laser during the shock heating. In the present experiments with large fractional absorption

and no significant noise problems, only one single scan of It behind the reflected shock

was analyzed to infer spectroscopic parameters.

The raw data traces of a typical experiment for high-temperature line-strength and

Ar-broadening measurements of CO are plotted in Figure 2.6(a). The laser intensity and

wavelength were scanned over the R(12) transition at 2190.02 cm−1 and recorded behind

the reflected shock at 1450 K and 1.63 atm (vibrational equilibrium) with a mixture of

0.496% CO/1% H2/Ar. Prior to each experiment, the shock tube is evacuated by a

turbomolecular pump and the baseline reference intensity (I0) recorded. The spectral

absorbance is then determined by the Beer-Lambert law and plotted as a function of

wavenumber calibrated using the etalon trace, as demonstrated in Figure 2.6(b). The

15

measurement was overlaid with a best-fit Voigt profile in the same figure. The peak-

normalized residual values are less than 0.8% over the entire absorption feature,

indicating that the Voigt profile adequately models the absorption line-shape.

(a) (b)

Figure 2.6 Illustration of (a) the measured raw-data traces (pressure, transmission through the shock tube

and the etalon) of the R(12) transition at 2190.02 cm−1, taken at 2.5 kHz with 0.496% CO/1% H2/Ar

mixtures behind the reflected shock (vibrationally equilibrated reflected shock conditions: 1450 K, 1.63

atm); (b) the reduced line-shape of the R(12) transition (solid line, top panel), its best-fit Voigt profile

(dashed line, top panel), and the residual (bottom panel).

The line-strength at a selected temperature can be inferred using (Eqn. 2-2) by

calculating the integrated absorbance of the target line from the best-fit Voigt values.

Figure 2.7 illustrates the measured line-strengths of four representative transitions at

1100-2000 K behind reflected shock waves. The calculated values using (Eqn. 2-5) with

the line-strength S(296K) and the lower-state energy E” from the HITRAN database [13]

are also plotted for comparison, illustrating excellent agreement (1- deviation between

1.4% and 1.8%) with our measurements.

Similarly, the collisional full-width at half maximum (FWHM) was inferred from

the Voigt fit of the absorption profile as shown in Figure 2.6(b). The collisional width is

dominantly affected by the Ar-broadening as CO (0.5%) and H2 (1%) are both

significantly diluted in argon. Thus, at a given temperature, the Ar-broadening coefficient

is inferred directly from the measured collisional width with self- and H2- broadening

neglected. Figure 2.8 plots the measured Ar-broadening coefficients (2CO-Ar) as a

function of temperature for CO transitions v” = 0, R(12) and R(13) and v” = 1, R(21). A

16

two-parameter best fit to the experimental data following (Eqn. 2-4) gives 2CO-Ar(296K)

and its temperature coefficient n, as illustrated in Figure 2.8 and summarized in Table 2.2.

Note that the errors quoted in the table for the experimental results correspond only to the

standard deviations derived by linear least-squared fits of log(2CO-Ar) versus log(296/T).

Experimental results of the line v” = 1, R(22) at 2194.46 cm-1 are not included in Table

2.2, since this line was found to be blended with a neighboring transition v” = 2, R(32) at

2194.44 cm-1 especially at higher temperatures, leading to larger uncertainties in the

measurement.

Figure 2.7 Comparison of the measured line-strengths for the CO transitions at high temperatures with the

HITRAN database [13].

Figure 2.8 Ar-broadening coefficient 2CO-Ar measurements for the CO transitions: R(12), R(13) and R(21).

The two-parameter best fit extrapolated to 296 K gives 2CO-Ar(296 K)=0.079±0.007 cm-1/atm and n =

0.581±0.012 for transition R(12), 2CO-Ar(296 K)=0.079±0.009 cm-1/atm and n = 0.600±0.016 for transition

R(13), and 2CO-Ar(296 K) = 0.072±0.007 cm-1/atm and n = 0.571±0.012 for transition R(21), respectively.

17

Table 2.2 Line-strength and broadening parameters for the CO transitions. Uncertainties of measurements

are given in the parentheses; the extrapolation of 2CO-Ar to 296 K following (Eqn. 2-4) is based on the

experimental data over the temperature range of 1100-2000 K.

Transition(v”, J”)

S @ 296 K,(cm-2atm-1)

2CO-Ar (296K),(cm-1/atm)

N

HITRAN Measured Bouanichet al. [31]

Measured@ 296 K

Fit to1100-2000 K

Fit to1100-2000 K

0, R(12) 7.13(2-3%)

7.16(2.3%)

0.088 0.088(3.0%)

0.079±0.007 0.581±0.012

1, R(21) 4.32×10-5

(2-3%)- - - 0.072±0.007 0.571±0.012

0, R(13) 6.04(2-3%)

5.95(2.3%)

0.087 0.085(2.9%)

0.079±0.009 0.600±0.016

0, P(20) 0.876(2-3%)

0.872(2.5%)

0.079 0.079(3.3%)

0.083±0.011 0.639±0.024

1, P(14) 2.64×10-4

(2-3%)- - - 0.074±0.018 0.560±0.045

In addition, the room-temperature (296 K) line-strengths and Ar-broadening

coefficients of the ground state transitions (v” = 0) can be directly determined by

examining a frequency scan prior to the passage of the incident shock. Figure 2.9

illustrates the variation of the measured integrated absorbance and FWHM with pressure

at 296 K for the representative transitions v” = 0, R(12) and v” = 0, P(20). Following

(Eqn. 2-2) and (Eqn. 2-3), the line-strength and Ar-broadening coefficient at 296 K are

inferred from the slope of the linear fit to the data as shown in Figure 2.9 (a) and (b),

respectively. These experimental results are also summarized in Table 2.2. The measured

line-strength at 296 K shows excellent agreement with the HITRAN database [13]

(within 1.5%), and the measured Ar-broadening coefficient is also in quite good

agreement with the previous room-temperature studies by Bouanich and Haeusler [31].

We also compared the 2CO-Ar(296 K) obtained from the direct measurements at

room temperature with the extrapolated values (assuming constant n) from the shock tube

measurements over the 1100-2000 K range. It should be noted that a 3-10% difference of

2CO-Ar(296 K) can be found between these two methods. This may be explained by the

fact that the temperature coefficient n in (Eqn. 2-4) itself is a weak function of

temperature over the range from 296 K to 2000 K. Since the Ar-broadening coefficient as

a function of temperature on a log-log plot is well-fit by a straight line as illustrated in

Figure 2.8, n can be treated as a constant over this specific temperature range of 1100-

18

2000 K and utilized in the sensor development for shock tube and combustion

applications.

(a) (b)

Figure 2.9 Room-temperature (296 K) spectroscopic parameter measurements for (a) line-strength using the

measured integrated absorbance versus P1 (20-60 Torr), and (b) Ar-broadening coefficient using the

measured collisional FWHM versus P1.

2.5 Sensor Validation in Shock Tube Experiments

CO concentration and temperature sensors using both scanned-wavelength and

fixed-wavelength direct absorption strategies are first validated in non-reactive shock-

heated gases before being used in combustion kinetics applications. The bandwidth of the

fixed-wavelength CO sensor is typically 1 MHz (limited by the detector bandwidth),

compared to 2.5 kHz for the scanned-wavelength scheme which is limited by the scan

rate of the laser injection current.

2.5.1 Scanned-Wavelength CO Sensor Using a Single QC Laser

Single-laser sensing has the advantages of simplifying the sensor system and

reducing cost. Transitions v” = 1, R(21) and v” = 0, R(12) with relatively large difference

in E” are close enough to be covered by a single scan of the QC laser. Temperature can

be inferred by comparing the measured peak absorbance ratio with the simulation. The

simulated peak absorbance ratio for this line pair is plotted in Figure 2.10 as a function of

temperature. Notice that the pressure effect is also investigated to show that the

19

uncertainty due to pressure variation is negligible in the pressure range of 1-2 atm. At

1500 K, for example, the temperature uncertainty is ~6 K (0.4%) with a pressure change

from 1 to 2 atm.

Figure 2.10 Simulated peak absorbance ratio for the line pair R(21)/R(12) and R(21)/P(20) using the

spectroscopic parameters listed in Table 2.2.

The experimental setup for the single-laser sensor validation in a shock tube is the

same as that shown in Figure 2.5. The test gas mixture is known to be 0.49% CO/ 2% H2

/Ar; similarly, hydrogen is added to accelerate vibrational relaxation. Figure 2.11

illustrates a representative laser absorption measurement of temperature behind the

reflected shock at 1526 K and 1.57 atm (vibrationally relaxed). The laser intensity and

wavelength were tuned across these two absorption profiles of interest at 2.5 kHz (top

panel in Figure 2.11), along with the corresponding absorbance profile shown in the

bottom panel. Assuming ideal shock conditions, the gas properties were reasonably

regarded to be unchanged within each scan of 0.4 ms. During the test time of 2.5 ms, the

sensor produced six data points of temperature as illustrated in the bottom panel of Figure

2.11, which were in good agreement (1527-1529 K in the first 1 ms, less than 0.2%

difference) with the known value calculated using normal shock equations. Notice that

the measured temperature drops significantly by ~30 K at 2.4 ms, possibly due to a weak

interaction of the reflected shock wave with the contact surface (driven and driver gas).

With the temperature measured, the CO mole fraction is then inferred from either

line of these two transitions. The CO mole fraction is measured to be (0.491±0.003)%

using line R(12), again showing good agreement with the known CO concentration of

20

0.49%. Moreover, at 2.4 ms when the non-ideal shock condition happens, the CO mole

fraction is still accurately measured to be 0.489%.

Experiments were repeated under different shock conditions to measure gas

temperature and CO concentration, and the results for a single scan are compared with

the known values and plotted in Figure 2.12 (solid squares). Good agreement can be seen

between measurements and the known values for both the temperature (1- deviation

0.8%) and CO concentration (1- deviation 1.6%).

Figure 2.11 Sample traces of laser transmission and pressure (top panel), as well as absorbance and

temperature (bottom panel) measured in non-reactive test gases (0.49% CO/ 2% H2 /Ar, vibrationally

equilibrated reflected shock conditions: 1526 K, 1.57 atm). A single QCL was used to scan over the line

pair R(21) and R(12) at 2.5 kHz.

Figure 2.12 Shock tube validation measurements for the scanned-wavelength (measured for a single scan

behind the reflected shock, solid squares) and the dual-laser fixed-wavelength (averaged over the first 0.3-1

ms after the shock, solid triangles) direct absorption CO sensors (0.49% CO/2% H2/Ar, 1.3-1.7 atm).

21

2.5.2 Fixed-Wavelength CO Sensor Using Two QC Lasers

The sensor bandwidth of scanned-wavelength direct absorption is limited to several

kHz, making it impossible to capture the rapid change of gas properties in chemical

reactions. Here a fixed-wavelength CO concentration and temperature sensor with a

bandwidth of ~1 MHz is developed for shock tube experiments using a dual-laser fixed-

wavelength diagnostic strategy.

Figure 2.13 is a schematic of the experimental setup. The light from each laser is

collimated and transmitted through a different pair of windows on the shock tube

sidewall. The laser wavelengths are fixed at the line-centers of the two selected

transitions v” = 1, R(21) at 2191.50 cm-1 and v” = 0, P(20) at 2059.91 cm-1, respectively.

This optical configuration utilizes the fact that the gas properties across the shock tube

are uniform.

Figure 2.13 Experimental setup for the fixed-wavelength two-line temperature and CO concentration

measurements in a shock tube.

Figure 2.14(a) demonstrates a sample time-history of the laser absorbance recorded

behind a reflected shock at 1454 K and 1.62 atm with 0.49% CO/2% H2/Ar mixture. The

laser absorption reaches the plateau level as the CO is fully relaxed at ~0.2 ms. Note that

the sensor essentially measures the vibrational temperature, and hence the absorbance in

the v” = 1, R(21) line increases with time, after the shock, from zero to its plateau value.

Conversely, the v” = 0, P(20) absorbance decreases from its elevated initial value to its

plateau value as vibrational relaxation takes place. Measured time-histories of pressure

and temperature are plotted in Figure 2.14(b). The average measured temperature over

the time interval 0.2-1.5 ms is 1456 K with a standard deviation of 6 K (0.4%), showing

excellent agreement with the calculated value of 1454 K. Note that the sensor is capable

2191.50 cm-1

Aperture

DetectorBP Filter

2059.91 cm-1

22

of capturing the slight rise of temperature at later times from about 0.9-1.8 ms, which

results from the non-ideal shock tube effects of boundary layer growth and incident-

shock attenuation. CO mole fraction can be computed using the measured pressure,

temperature and transmission signal of either laser.

The experimental results are summarized and plotted in Figure 2.12 (solid triangles).

The measured and calculated temperatures are in good agreement (within 1.1%) over the

tested temperature range of 1200-1900 K, and the measured mole fraction agrees with the

known values within 1.7%. These results confirm the sensor accuracy for temperature

and CO concentration measurements at combustion temperatures. It is noteworthy that

this sensor has the potential to measure both translational/rotational temperatures and

vibrational temperatures, when these temperatures differ, by employing line pairs with

the same v” value or with different v” values, respectively.

(a) (b)

Figure 2.14 Fixed-wavelength temperature measurements using two QC lasers with 0.49% CO/2% H2/Ar:

(a) measured absorbance traces for the two lasers; (b) measured temperature and pressure. Vibrationally

equilibrated reflected shock conditions: P5 = 1454 K, T5 = 1.62 atm.

23

2.6 Temperature and CO Concentration Measurements in

Combustion Gases

Shock tubes are used to study gas phase combustion reactions by measuring ignition

delay times and by monitoring species time-histories over a wide range of temperatures

and pressures [5]. Accurate, time-resolved measurements of combustion species in shock

tubes are therefore critical, with laser absorption the most commonly employed method

[5,32]. Here the fixed-wavelength CO sensor validated in Section 2.5.2 is demonstrated

in a kinetic study of the high-temperature pyrolysis and oxidation of methyl formate

(MF), a simple biodiesel surrogate.

Figure 2.15 Absorbance time-histories of R(21) and P(20) during the pyrolysis of methyl formate. Initial

reflected shock conditions: T5 = 1364 K, P5 = 1.63 atm, 0.5% MF/Ar.

The shock tube/laser diagnostic experimental setup is the same as that shown in

Figure 2.13. Typical absorbance time-histories for both lasers are plotted in Figure 2.15

for a test mixture of 0.5% MF/Ar shock-heated to 1364 K, 1.63 atm. The absorbance

profile for each laser directly represents the CO formation during the high-temperature

pyrolysis of methyl formate. Temperature is inferred from the measured absorbance ratio,

showing the slight decrease (15 K) that occurs during the decomposition of methyl

formate; see Figure 2.16(a). Our measurement is compared with a chemical kinetic

simulation using the Dooley et al. [33] mechanism, performed in a commercial software

package CHEMKIN-PRO [34]. As illustrated in Figure 2.16, the simulation results are

insensitive to the selection of different gasdynamic constraints, e.g. constant volume (V)

and internal energy (U), or specified pressure (P) and enthalpy (H).

24

(a) (b)

Figure 2.16 Temperature and CO concentration measured during a shock with initial mixture of 0.5%

MF/Ar; simulations using the Dooley et al. [33] mechanism are shown for comparison. Initial reflected

shock conditions: T5 = 1364 K, P5 = 1.63 atm.

(a) (b)

Figure 2.17 Temperature and CO concentration measurements during MF oxidation for a mixture of 0.494%

MF, 0.988% O2 (= 1) and Ar; simulations using the Dooley et al. [33] mechanism are shown for

comparison. Initial temperature and pressure behind the reflected shock are T5 = 1379 K, P5 = 1.67 atm.

CO concentration time-history is then inferred from either absorption trace; here the

R(21) transition is used. Figure 2.16(b) compares the sensor measurement with the

simulation using the Dooley et al. [33] mechanism. The model underpredicts the rate of

CO formation by about 10% during MF pyrolysis, revealing the need for minor

modifications to the kinetic model. The difference between simulations under constant U,

V and specified H, P constraints is negligible.

Similar measurements were performed for high-temperature MF oxidation in the

shock tube. The measured temperature and CO mole fraction are plotted in Figure 2.17

25

for a shock with 0.494% MF and 0.988% O2 (= 1) in Ar as the initial mixture. The

measured temperature time-history shown in Figure 2.17(a) reveals that the gas

temperature remains almost constant before the ignition happens at ~1 ms, and then rises

significantly by 340 K at 1.5 ms. This significant temperature change is caused by heat

release due to MF ignition at the time ~1 ms. Unfortunately, there is no accepted way to

simulate this post-ignition process (except for very dilute mixture) as it is not a

homogeneous reactor with a simple gasdynamic constraint, e.g. constant U, V. Hence,

modeling the temperature and species time-histories are typically done only prior to the

ignition event.

Figure 2.17(a) compares the measured temperature time-history with simulations

under constant U, V and specified H, P constraints. The simulation results strongly

depend on the selection of gasdynamic constraints due to the large amount of heat release

after ignition. Since the temperature rises significantly (from 1379 K to 1719 K) during

the MF oxidation, it is critical to take into account these temperature and pressure

changes in specifying the absorption coefficient when inferring the CO mole fraction.

Figure 2.17(b) illustrates a comparison of the uncorrected CO concentration (assumes

unchanged temperature and pressure) with the corrected values using the measured

temperature and pressure data. A modest discrepancy (5.3% difference) is seen after 1 ms

when ignition starts in the reaction system. Reasonable comparison can be made between

measurements and simulations only prior to MF ignition. Interestingly, the simulations

under constant U, V and specified H, P constraints predict almost the same early-time CO

formation till 0.9 ms but differ significantly after that. The simulations using Dooley et al.

[33] show good agreement with our measurements at early times (<0.2 ms) and

accurately predict the peak value of CO (8420 ppm in experiment, compared to 8560

ppm in simulation) before starting to decline at 1 ms. Quantitative data sets such as these

should greatly aid the validation of existing kinetic mechanisms. In addition, a recent

reported constrained-reaction-volume strategy by Hanson et al. [35] can effectively

eliminate or minimize pressure changes due to combustion heat release, enabling

quantitative modeling of the kinetics throughout the combustion event using a simple

assumption of specified pressure and enthalpy.

26

27

Chapter 3. Two-Line Thermometry for

Multiphase Combustion Flows

3.1 Introduction

Accurate knowledge of temperature is very important in combustion studies of

chemical reaction rates, process efficiency and pollutant emissions [36,37]. Shock tubes

are typically used to study chemical kinetics at elevated temperatures since they provide a

well-controlled pressure and temperature environment. Currently there is a need to

investigate the combustion chemistry of real fuel blends, including diesels, jet fuels and

biodiesels, all of which have low vapor pressures precluding their study in conventional

unheated shock tube kinetics experiments. Thus, we have designed a new type of shock

tube [38], in which a spatially-uniform fuel aerosol is loaded into the shock tube; the

evaporation of this aerosol behind the incident shock wave is used to produce fuel vapor

for the subsequent reaction behind the reflected shock wave. For a better understanding

of the shock tube performance, accurate sensors with rapid time-response are now

required to explore and validate the test conditions in the evaporating fuel aerosol.

CO2 is a particularly attractive target species since it is a primary combustion

product of hydrocarbon fuels and can be added as an inert tracer for the measurements in

both non-reactive and many reactive flow environments. Sensors for CO2 previously used

three different vibrational combination bands, near 1.57 m (2v1+2v2+v3) [9,10,18], 2.0

m (v1+2v2+v3) [7,20,39], and 2.7 m (v1+v3, 2v2+v3) [40,41]. The combination bands

near 2.7 m offer the most promising candidate transitions in terms of their stronger

absorption (approximately 1000 and 50 times stronger, respectively, in contrast to the

combination bands near 1.57 m and 2.0 m). The first CO2 concentration and

28

temperature sensor for combustion gases using diode laser absorption in the v1+v3 band

has recently been reported by Farooq et al. [40].

To our knowledge, few studies of temperature measurements in evaporating aerosol

have been conducted using optical diagnostics, due to the problem of dealing with the

interference of droplet extinction. Beyrau et al. reported an application of pure rotational

coherent anti-Stokes Raman spectroscopy (CARS) in quantitative gas-phase temperature

measurements in the vaporizing spray of an automotive fuel injector [42]. Awtry et al.

developed a TDL spectrometer based on scanned-wavelength direct absorption (DA) for

multi-species measurements in a dense water mist environment [43]. Porter et al. used a

three-wavelength mid-IR absorption/extinction diagnostic to measure the temperature and

concentration of n-decane using transitions in the C-H stretching band near 3.4 m [44].

However, these mid-IR measurements with broad absorption features typically require

precise knowledge of the aerosol droplet properties, together with an additional extinction

measurement with a non-resonant beam.

In this work, we extended the 2.7 m fixed-center-wavelength WMS-2f sensor of

Farooq et al. [40] to the sensitive and accurate temperature measurements of CO2 in an

evaporating aerosol. The sensor was used for temperature measurements in shock-heated

n-dodecane aerosols with CO2 as an inert tracer. The temperature decline immediately

behind the incident shock due to aerosol evaporation was successfully captured,

illustrating a good agreement with modeled values. Measurement uncertainties resulting

from the pressure fluctuation and concentration change were investigated to confirm that

the two-line thermometry was sensitive only to temperature under the conditions studied.

3.2 Wavelength Modulation Spectroscopy Fundamentals

The quantitative measurement of gas properties requires an accurate WMS model.

The model used in this work is based on Li et al. [45], which includes actual laser

performance parameters. In tunable diode laser (TDL) WMS, the laser injection current is

sinusoidally modulated with an angular frequency of = 2f to produce frequency (FM)

and intensity modulation (IM) as

( ) cos( )v t v a t , Eqn. 3-1

29

00 0 1 2 2( ) [1 cos( ) cos(2 )]I t I i t i t , Eqn. 3-2

where v [cm-1] is the center laser frequency; a[cm-1] is the modulation depth; i0 and i2 are

the linear and nonlinear IM amplitudes normalized bythe average laser intensityI0, and

and are the linear and nonlinear FM/IM phase shifts. The transmission coefficient

(v) for the laser light through the absorbing gas medium (assumed uniform) of length

L[cm] is described by Beer’s law

0

exptv

v

Iv

I

, Eqn. 3-3

where It is the transmitted laser intensity. As discussed before, v is known as the spectral

absorbance

v i vP S T L , Eqn. 3-4

where P[atm] is the total gas pressure; i is the mole fraction of the absorbing species,

and S(T)[cm-2/atm] and v[cm] are the line-strength at temperature T[K] and the line-

shape function of the absorption feature, respectively.

The time-dependent transmission coefficient (v) can be expanded in terms of a

Fourier series with ,kH v a the kth-order Fourier component as discussed in [45]. The

case of k = 2 is generally of highest interest in WMS, since it is closely related to the

absorption. The 2f signal can be experimentally demodulated from the detector signal

using a lock-in amplifier, and the X and Y components can be expressed as

0 0 42 2 1 3 1 2 0 2cos cos

2 2 2f

i HGIX H H H i H

, Eqn. 3-5

0 0 42 1 3 1 2 0 2sin sin

2 2 2f

i HGIY H H i H

, Eqn. 3-6

where the magnitude of WMS-2f signal is given by = + . G accounts for

the optical-electrical gain of the detection system, and also transmission losses other than

absorption. This factor can be removed by normalization with the 1f signal, which can be

calculated:

30

20 2 2

1 1 0 0 1 1 3 2

1/ 22

2 20 0 1 1 3 2

cos cos2 2 2

sin sin .2 2

f

H iGIR H i H H H

H ii H H H

Eqn. 3-7

Gas temperature is inferred from the ratio of the two 1f-normalized WMS-2f signals/ / , and is closely related to the ratio of absorption line-

strengths.

In many practical combustion processes, fuels are injected as a liquid spray which

quickly evaporates at elevated temperatures. Expanding Beer’s law by incorporating

extinction due to scattering by aerosol particles, the laser light extinction coefficient can

be expressed as [46]:

0

lnv v v

v

IExt k L

I

,Eqn. 3-8

where kv[cm-1] is the spectral absorption coefficientof the gas and v[cm-1] is known as

the droplet extinction coefficient. From models of Mie scattering [46], the extinction

coefficient v is an integration of droplet parameters over the full range of droplet

diameter, D, given by

max

min

2

, ,4

D

v ext v

D

DN f D Q D m dD

,Eqn. 3-9

where N[cm-3] is the droplet loading; f(D)[cm-1] is the droplet size distribution function;

Qext,v is the normalized extinction cross section, and m is the complex refractive index.

In the aerosol shock tube experiments described later, the performance of the WMS

temperature sensor was evaluated in the presence of evaporating n-dodecane aerosols.

Since the two wavelengths selected for the temperature sensing of CO2 are both near 2.7

m, the absorber CO2 has sharp absorption structures as illustrated in Figure 3.1. Using a

readily available Mie scattering code [47], the droplet extinction cross section Qext,v was

calculated and shown in Figure 3.2 for two different median droplet diameters. For the

two selected sensor wavelengths, the droplet extinction term can be treated as constant.

Hence, the droplet extinction reduces the magnitudes of the 2f and 1f signals equally and

31

1f-normalization of the 2f signal provides an effective means of excluding the effects of

droplet scattering on accurate temperature measurements.

Figure 3.1 Calculated CO2 (1%) absorption spectra for the R(28) transition at 3633.08 cm-1 (2752.48 nm)

and the P(70) transition at 3645.56 cm-1 (2743.06 nm) under typical shock tube conditions: T2 = 650 K, P2

= 0.5 atm; T5 = 1200 K, P5 = 1.0 atm; L = 10 cm.

Figure 3.2 Calculated extinction cross section (Mie scattering code [47]) for liquid n-dodecane droplets; Dm

is the median diameter of aerosol droplet size.

Table 3.1 Measured spectroscopic data for the selected CO2 line pair (from [40]).

2600 2800 3000 3200 3400 3600 38001.0

2.0

3.0

4.0

5.0

Frequency [cm-1]

Qex

t

Dm= 1.5 m Dm= 2.5m

v0

[cm-1]

E”[cm-1]

S(296K)

[cm-2/atm]

self(296K)

[cm-1/atm]

nself Ar(296K)

[cm-1/atm]

nAr

3633.08 316.77 6.13E-01 0.171 0.654 0.112 0.658

3645.56 1936.09 7.04E-04 0.130 0.695 0.091 0.694

32

3.3 Sensor Design

3.3.1 Line Selection

The selection of an optimum CO2 line pair was the first step in the TDL sensor

design. Two CO2 absorption transitions in the v1+v3 combination band were selected for

the WMS-2f/1f sensor under the conditions studied. The selected transitions R(28) and

P(70) have line-centers near 3633.08 cm-1 (E" = 316.77 cm-1) and 3645.56 cm-1 (E" =

1936.09 cm-1), respectively. The well-separated values for lower-state-energy E’’ enable

high temperature sensitivity in the range 600-1600 K. Spectroscopic parameters for the

two transitions have been measured previously in a heated static cell [40] and compared

to the HITRAN database [13]. Figure 3.1 depicts the simulated absorption spectra for 1%

CO2 in argon and the two lines selected, using the spectroscopic data in Table 3.1. These

simulations were performed with Voigt line-shape functions for typical shock tube

conditions: incident shock (T2 ~650 K, P2 ~0.5 atm) and reflected shock (T5 ~1200 K, P5

~1.0 atm).

3.3.2 Measurement Uncertainties

Gas temperature was determined by comparing the measured WMS-2f/1f signal

ratio to the simulation at the specified pressure and gas concentration. Note that the 2f

peak height ratio is not strictly a function of temperature alone, but also includes

contributions from pressure and gas composition through the line-shape function [48].

This signal variation was considered in evaluating measurement errors.

The influence of pressure variation on the 2f ratio is illustrated in Figure 3.3 for

representative reflected shock conditions. Although the 2f signal magnitude for each

individual line varies with the pressure, the 2f/1f ratio of the line pair is nearly

independent of pressure as shown in Figure 3.3(b). At 1200 K, a pressure change from

1.0 to 1.5 atm would result in 1% change in the peak ratio, corresponding to a

temperature change of only ~4 K (within 0.3%).

The influence of gas composition on the inferred temperature was also investigated.

For the non-reactive shock-heated CO2/Ar mixtures studied, the mole fraction of CO2

was known to be 2%. However, when fuel aerosol was also introduced, the CO2/Ar gas

33

mixture was diluted and the CO2 concentration was not exactly 2%. Figure 3.4 reveals the

fact that a 10% change in CO2 mole fraction generates a negligible temperature error of

less than 1 K for an incident shock condition of T2 = 650 K, P2 = 0.5 atm. Figure 3.4(b)

also illustrates the case when the test gas is optically thick (20% CO2 in argon,

dramatically different from the nominal concentration); even for this extreme case, the

error in the inferred temperature is only ~50 K, which obviously could be improved by

the simple iteration of CO2 temperature and concentration to obtain the accurate

temperature [48].

Figure 3.3 WMS simulation for (a) 2f/1f magnitude of 2752nm line and (b) 2f/1f ratio of 2752nm /2743nm

line pair, as a function of temperature for specified pressures at optimized modulation depths; 2% CO2 in

Ar, T = 900-1600 K, L = 10 cm.

Figure 3.4 WMS simulation for (a) 2f/1f magnitude of 2752nm line and (b) 2f/1f ratio of 2752/2743 line

pair, as a function of temperature for specified CO2 concentrations; P = 0.5 atm, T = 400-900 K, L = 10 cm.

34

3.4 Temperature Measurement in CO2/Ar Gas

3.4.1 Experimental Setup

(a) Top view

(b) Side view

Figure 3.5 Shock tube experimental setup.

Experiments were first performed in a pressure-driven shock tube with a purely

gaseous system; see reference [49] for further details of the 14.1 cm diameter shock tube.

A schematic of the experimental setup for two-wavelength measurements on the shock

tube is described in Figure 3.5. Two diode lasers from Nanoplus were sinusoidally

modulated by 100 kHz digital waveforms, for the selected transitions R(28) near 3633.08

2752nm 2743n

m

Shock tube

Wave

meter

Laser

controllers

DA

Q

Detector

Pressure transducer

100kHz

2752 nm 2743 nm

Trigger

Filter

2 cm

70 cm

Different test locations for

incident and reflected shocks

35

cm-1 (14.5°C, 80.0 mA) and P(70) near 3645.56 cm-1 (26.7°C, 143.0 mA). The

modulation depths were adjusted to the optimal values where the modulation index m =

2a/v = 2.2 [41], where v is the FWHM of the absorption line-shape. In these

measurements, we used a = 0.067 cm-1 and 0.056 cm-1 for transitions R(28) and P(70),

respectively, to achieve sensitive detection under the experimental conditions. On the

collection side the collimated beam from each laser was focused onto a liquid-nitrogen-

cooled InSb detector (IR Associates IS-2.0). The detector signals were sampled at a rate

of 10 MHz and demodulated by a digital lock-in amplifier on LabVIEW with a low-pass

filter bandwidth of 40 kHz to extract the 1f and 2f signals. Prior to each experiment, the

shock tube was first evacuated by a turbomolecular pump and the background signals

were recorded for both lasers. The background signals account for the ambient absorption

and the nonlinear intensity modulation in the laser and must be vector-subtracted from

the absorption signals [48].

3.4.2 Experimental Results

The initial gas temperature and pressure behind the reflected shock are accurately

known, providing a well-controlled environment to test the TDL sensor. The laser

diagnostic was typically located 2 cm from the endwall. A measured time-history of

temperature behind a reflected shock with a 2% CO2/Ar mixture initially at P1 = 55 Torr,

T1 = 298 K is plotted in Figure 3.6. The pressure recorded with a Kistler transducer is

also plotted. The average measured temperature over the initial time interval 0.1-0.6 ms

was 1193 K with a standard deviation of ~5 K, which was in excellent agreement (within

1%) with the expected value of 1199 K calculated using normal-shock relations. The

sensor successfully captured the slight rise of temperature (beginning at ~0.6 ms, with

T/T = +1.6%) behind the shock wave which may be attributed to the effects of

boundary layer growth and shock attenuation. The time-resolved temperature trace in

Figure 3.6 indicates that the rarefaction wave arrives at ~1.5 ms, after which the

temperature and pressure decline. Additional experiments were conducted under different

shock conditions, and Figure 3.7 (square points) compares the measured temperatures

(averaged over the time interval 0.1-0.5 ms) with the expected values. These comparisons

confirm good agreement (within 1.5%) over the entire 1100-1500 K temperature range.

36

Figure 3.6 Measured temperature and pressure trace during a shock with CO2/Ar mixture without aerosol.

Initial conditions: P1 = 55.0 Torr, T1 = 298 K; incident shock (calculated): P2 = 0.43 atm, T2 = 697 K;

reflected shock (calculated): P5 = 1.48 atm, T5 = 1199 K.

Figure 3.7 Temperatures measured by the WMS-2f/1f sensor in shock-heated CO2/Ar mixture without

aerosol versus calculated values using shock jump equations; ±1.5% error bars. The square points represent

T5 behind reflected shocks (P5 = 1.0-1.5 atm); the triangular points represent T2 behind incident shocks (P2

= 0.4-0.6 atm).

The time interval between the arrival of the incident and the reflected shock was

only 60-70 s at 2 cm from the endwall, and the time-history of T2 was not captured by

the sensor. Thus, a second test location was established 70 cm from the endwall to

measure T2 versus time, and the measured temperatures immediately following the shock

waves were also compared with calculations as shown in Figure 3.7 (triangular points,

±1.5% error bar). The agreement between the measured and the known temperature was

37

very good, ~1.5% over the full 650-1500 K range. Figure 3.8 illustrates the temperature

versus time for an incident shock wave with a 0.35 atm, 649 K post-shock condition. The

measured temperature averaged within the plateau region of ~2.7 ms was 645 K, within

0.6% of the frozen-chemistry shock calculation.

Figure 3.8 Measured temperature and pressure trace at 70 cm from the endwall with CO2/Ar mixture. Initial

conditions: P1 = 50.1 Torr, T1 = 298 K; incident shock (calculated): P2 = 0.35 atm, T2 = 649 K.

3.4.3 Comparison of CO and CO2 Thermometry

CO and CO2 are both attractive target species for gas temperature sensing in

combustion. Current CO two-line thermometry has been developed using absorption

transitions in the fundamental bands of CO between 4.5 and 4.8 m. It provides hundreds

of times stronger absorption compared to the current CO2 temperature sensor using the

combinational bands near 2.7 m. However, CO is a reactive species at high

temperatures and can cause unwanted interfering reactions in shock tube chemical

kinetics studies. The vibrational relaxation time of 0.5% CO in Ar is estimated to be 5 ms

at 1200 K and 1 atm [30], which is beyond the normal test time of 2 ms in a shock tube.

In order to reduce the CO relaxation time to be within 0.1 ms, more than 10% helium

(inert gas) needs to be added to accelerate CO vibrational relaxation. Hence, the

additional information of CO/He collisional broadening coefficients is required for

accurate species and temperature sensing.

CO2 two-line thermometry seems to be more promising for temperature

measurements in shock tube experiments. The seeded small amount of CO2 would not

38

significantly affect the chemical kinetics and the CO2 vibrational relaxation happens

within tens of microseconds behind reflected shock waves. In order to increase the CO2

detectivity, much stronger absorption transitions in the fundamental bands near 4.3 m

are suggested for future work.

3.5 Sensor Validation in a Aerosol Flow Cell

After the temperature sensor was validated for gas-phase measurement, we

investigated its performance in the presence of aerosol scattering. A series of

measurements with different aerosol loadings were made in a flow cell described in

Figure 3.9. The n-dodecane aerosol (approximately log-normal droplet size distribution

with median diameter of ~3 m) was produced by an ultrasonic nebulizer, entrained in a

flow of CO2/Ar mixture, and passed through the 5.8 cm path-length cell. Aerosol loading

was varied by altering the flow rate over the nebulizer, and the pressure inside the cell

was monitored by a pressure transducer and maintained at 1 atm during the measurement.

Since the aerosol loading is proportional to the droplet extinction coefficient v [46], we

use off-line extinction to indicate the level of aerosol loading, by tuning the laser

wavelength off the CO2 absorption transition.

Figure 3.9 Aerosol flow cell experimental setup.

The P(70) transition near 3645.56 cm-1 (E’’ = 1936.09 cm-1) was too weak to be

detected in the flow cell designed for experiments at room temperature. Therefore, only

the individual line R(28) of the line pair was studied to demonstrate that the WMS-2f/1f

signal with droplets present in the gas flow was the same as that in the gas-phase mixture.

39

In contrast to the fixed-center-wavelength WMS used in shock tube measurements, a

scanned-wavelength WMS was utilized here to obtain the complete 2f line-shape so that

the 1f-normalization accounting for aerosol extinction was clearly observed near the line-

center. For this purpose, in addition to the high-frequency sinusoidal modulation on the

laser injection current, a repetitive linear ramp was superimposed on the modulation

current to sweep the laser wavelength across the absorption feature.

Figure 3.10 illustrates a typical example of measured WMS- 2f, 1f and 2f/1f data at

different aerosol loadings (indicated by the droplet extinction coefficient v) for the CO2

transition R(28) near 3633.08 cm-1 at P = 770 Torr, T = 303 K. Prior to the injection of

the aerosol, the 2f and 1f signals for the spectral absorption of 2% CO2 in argon were

recorded; their line-shapes were plotted as solid lines in Figure 3.10 (a) and (b),

respectively, reflecting the fact that the kth Fourier amplitude is proportional to the kth

derivative of the absorbance [45]. The absolute 2f peak was 0.083 with the corresponding

1f signal of 0.276, which resulted in the normalized peak value 2f/1f = 0.301 (0.2%

difference from the simulation). After the n-dodecane aerosol was loaded into the flow

cell, the 2f and 1f signals still maintained the same line-shapes (see Figure 3.10), but the

absolute amplitude decreased with larger aerosol loadings. This deviation was

successfully eliminated by the 1f-normalization strategy as shown in Figure 3.10(c),

where the measured 2f/1f peaks were almost constant (0.301±0.004) despite the

interference of droplet extinction.

With the increase in aerosol loading, the transmitted laser intensity decreased due to

the increased droplet scattering. The performance of this WMS-2f/1f sensor was

evaluated over a wide range of droplet extinction (0-99.5%). Figure 3.11 compares the

measured WMS-2f/1f data with the gas-phase data simulation which ignored the droplet

scattering. The 1f-normalized WMS-2f sensor provided an error of <2% in the 2f/1f

signal (corresponding to 1.6% in temperature) for transmission losses by droplet

scattering as large as 99.5%.

40

Figure 3.10 Measured WMS- (a) 2f, (b) 1f and (c) 2f/1f signals in an aerosol flow cell for the R(28)

transition of CO2 with different aerosol loadings; v represents the droplet extinction coefficient.

Figure 3.11 Comparison of the measured WMS-2f/1f data with the simulated value under the condition of

no droplet scattering.

41

3.6 Temperature Measurement in Shock-Heated Aerosol

Next, the sensor was demonstrated for accurate temperature measurements in the

aerosol shock tube. This facility was developed in our laboratory to conduct studies of

droplet evaporation kinetics and subsequent chemical reactions of the fuel vapor [38].

The gas/aerosol mixture was prepared in a separate holding tank and then carefully drawn

into the shock tube (with an inner diameter of 10 cm) by a slightly under-pressure dump

tank. The aerosol was rapidly evaporated by the incident shock-heating, and the resulting

vapor-phase mixture was shock-heated a second time by the reflected shock wave. In the

present experiments, the incident shock heated the aerosol in the test gas mixture to 520-

700 K, vaporizing the fuel droplets, and the WMS-2f/1f sensor was used to capture the

transient temperature variation during droplet evaporation.

As discussed in Section 3.4.1, all of the experimental procedures were nearly

identical, except for the fact that the micron-sized aerosol of n-dodecane liquids was

produced by a nebulizer and mixed with the CO2/Ar test gas. The test location was set at

10 cm from the endwall to obtain adequate test time to observe evaporation behind the

incident shock as well as to obtain an accurate estimation of the temperature behind the

reflected shock.

Figure 3.12 Measured temperature for an incident shock-heated aerosol with the WMS-2f/1f sensor located

at 10 cm from the endwall: (P2)w/o evap = 0.50 atm, (T2)w/o evap = 558 K; (P2)post evap = 0.54 atm, (T2)post evap =

528 K; P5 = 1.79 atm, T5 = 796 K. A non-resonant 660 nm laser is used to indicate the droplet scattering.

42

Figure 3.13 Temperatures measured in aerosol shock tube by the WMS-2f/1f sensor versus calculated

values using numerical code; ±1.8% error bars. The square points represent T5 behind reflected shocks (P5

= 1.0-1.5 atm); the triangular points represent the post-evaporation T2 behind incident shocks (P2 = 0.4-0.6

atm).

A sample measurement is shown in Figure 3.12 that plots the time-history of the gas

temperature in evaporating n-dodecane aerosold determined from the WMS-2f/1f sensor,

and the transmission trace of a visible (660 nm) laser beam as indicated by the solid line.

Time zero represents the arrival of the incident shock wave, after which the droplets

begin to evaporate. The light extinction (defined in Eqn. 3-5) of the visible beam

indicates that there was a significant attenuation by droplet scattering before the aerosol

was fully evaporated near 0.4 ms. Accordingly, with the evaporative cooling of the

droplets, the measured gas temperature decreased in the post-shock region until the

droplets evaporated completely as shown in the temperature time-history. The

measurement was compared with a calculation using a numerical code AEROFROSH,

developed in-house and based on ideal shock equations and thermodynamic conservation

conditions [38,50]. The results are plotted as dashed lines in Figure 3.12. The upper

temperature limit (T2)w/o evap=558 K was the initial temperature of the test mixture prior to

evaporation and the lower limit (T2)post evap=528 K was the temperature after the aerosol

evaporated completely. The sensor successfully provided a measurement of the

temperature change due to the evaporation of aerosol. After the arrival of the reflected

shock at 0.8 ms, the vapor mixture was heated and compressed a second time and the

measurement again was in good agreement (within 1.5%) with the calculated temperature.

43

Similar measurements were repeated with different shock conditions, as illustrated in

Figure 3.13, and the measured temperatures were within 1.8% of the calculated values

over the 520-1200 K range.

Calculations of the gas pressure and temperature behind shock waves assumed

vibrational equilibrium. Vibrational relaxation in the lower energy levels of CO2 occurs

rapidly behind these shock waves in CO2/Ar hydrocarbon mixtures; in fact, measured

vibrational relaxation time under our experimental conditions have been shown to be less

than 20 s [51]. It is possible that the cold boundary layers at each shock tube wall may

influence the path-integrated temperature along the beam path, causing measurement

uncertainties. We estimated the maximum possible thickness of the boundary layer at ~1

mm at the typical pressure and temperature in this work [52]. Assuming a 2 mm path in a

cold boundary at 300 K, we calculated the change in the absorbance ratio to be <1%,

implying a maximum temperature uncertainty of 3 K for T2 = 600 K and 3.9 K for T5 =

1000 K. Thus, the boundary layers are expected to have a negligible impact on

temperature measurements in the current shock tube, and the measured temperature time-

histories accurately describe the core flow behavior.

44

45

Chapter 4. Thermal Decomposition of

Methanol and Methyl Formate

4.1 Introduction

Alcohol fuels are recognized as promising renewable energy resources and are also

used as additives in gasoline to reduce the formation of poly-aromatic hydrocarbon

compounds, particulates, and soot [53,54]. Combustion studies of methanol, which shares

many chemical kinetic characteristics with higher alcohols, can shed light on the

combustion chemistry of alcohols in general.

Biodiesel, typically derived from a variety of vegetable oils, animal fats, and algae, is

one of the sustainable alternatives to fossil fuels [54,55]. It is an oxygenated, diesel-like

fuel consisting primarily of fatty acid methyl esters (FAMEs). Methyl formate (MF,

CH3OCHO) is the simplest methyl ester, and its study assists in understanding the effects

of oxygenated chemical structure that are characteristic of biodiesel fuels on reactivity

and pollutant formation. A fundamental study of MF kinetics is thus of immediate

interest to fuel modelers. Equally important, the reactions of methanol and methyl

formate comprise important and fundamental subsets of detailed hydrocarbon combustion

mechanisms [56,57]. Therefore, a thorough understanding of the combustion chemistry

of these basic fuels is relevant for constructing kinetic models of larger oxygenated and

hydrocarbon fuels.

Experimental investigations providing species time-history data describing methanol

and methyl formate combustion chemistry are particularly needed. Laser absorption

diagnostics, due to their fast time response and non-intrusive, in situ capability, are being

utilized increasingly in chemical kinetic studies [5], and can be used to directly measure

46

species concentration time-histories in shock tube experiments. These species data are

critically important to efforts aimed at validating large reaction mechanisms and refining

their component sub-mechanisms. The main purpose of this work is to improve

understanding of the pyrolysis of methanol and methyl formate through multiple species

time-history measurements and to identify areas within the kinetic models where

improvements are necessary.

4.2 Experimental

All experiments were performed in the same shock tube with a 15.24 cm inner

diameter as discussed before. Between experiments, the shock tube driven section and

mixing manifold were turbo-pumped at least 30 minutes, down to ~6 torr to remove

residual impurities. Research grade high-purity argon (99.999% pure, Praxair Inc.) was

used without further purification. Methanol and methyl formate (>99% pure, Sigma-

Aldrich) were frozen and degassed three times to remove dissolved volatiles before

making the mixtures. All the test mixtures were manometrically prepared in a stainless-

steel mixing tank (40 L) heated uniformly to 50°C with an internal magnetically driven

stirrer. Laser absorption and side-wall pressure measurements (Kistler 601B1 PZT) were

located 2 cm from the shock tube end wall. In this study, two laser absorption diagnostics

are utilized for accurate, time-resolved measurements of CH3OH and CO concentration

time-histories.

4.2.1 QC Laser Absorption of CO at 4.56 m

Absorption measurements of CO were made using the same quantum cascade laser

(QCL) as introduced in Chapter 2. A fixed-wavelength direct-absorption strategy was

employed in the present study to monitor the peak intensity of the R(13) absorption line

at 2193.36 cm-1. The spectroscopic parameters for the R(13) transition, including the line-

strength and self-broadening coefficient, were taken directly from the HITRAN database.

The collisional broadening coefficient for CO with argon (not available in HITRAN) was

measured in shock tube experiments over the temperature range of 1000–1800 K; results

can be found in Chapter 2.

47

4.2.2 CO2 Laser Absorption of Methanol and Methyl Formate

Methanol is monitored in shock tube experiments using continuous wave (cw) CO2

laser absorption at 9.676 m (1033.5 cm-1). This absorption diagnostic takes advantage of

the strong overlap of the P34 CO2 laser line, associated with the (0 0 1) to (0 2 0)

vibrational transition, with the strong Q-branch of the v8 methanol band [58]. We utilized

a grating-tuned CO2 gas laser (Model Lasy-4G, Access Laser Co.) with ~100 mW output;

a schematic of the typical shock tube experimental setup is described in Appendix A. The

methanol cross-section data were measured at 665-1014 K and 0.4-0.8 atm behind the

incident shock waves and at 1126-1940 K and 1.4-2.7 atm behind the reflected shock

waves; see Appendix A for details of the diagnostic scheme.

Methanol is one of the main products during MF decomposition, but interfering

absorption from MF at the early times exists at this wavelength. To separate these two

signals, absorption measurements were made at two different wavelengths, 9.67 m and

9.23 m, in repeated shock wave experiments at near-identical conditions. With separate

calibrations of the different absorption cross-sections for the two species at the two

wavelengths, the time histories for both MF and methanol can be inferred from the two

measured absorbance time-histories; see Appendix A. Uncertainties in the absolute

methanol or methyl formate concentration are typically 5%, dominated by uncertainties

in the absorption cross-section measurements.

4.3 High-Temperature Methanol Pyrolysis

All the exact experimental conditions are summarized in Table 4.1. Methanol

pyrolysis was studied behind reflected shock waves using methanol and CO absorption

diagnostics at 9.676 m and 4.56 m, respectively. Figure 4.1 compares measured

methanol and CO concentration time-histories at different temperatures with those

predicted from the Li et al. [17] mechanism under the standard constant energy (U) and

volume (V) constraints using CHEMKIN PRO [34] software package. Here we have

chosen the Li et al. [17] mechanism as the base mechanism in the following analysis,

which is a detailed kinetic model optimized for methanol combustion.

48

Table 4.1 Summary of current methanol and methyl formate pyrolysis experiments.

Methanol PyrolysisXmethanol = 1% Xmethanol = 0.2%

T5 (K) P5 (atm) T5 (K) P5 (atm)1266 2.5 1403 1.21368 2.4 1507 1.11458 2.3 1623 1.11567 2.1 1707 1.31610 2.2

Methyl Formate PyrolysisXmethyl formate = 1% Xmethyl formate = 0.1%

T5 (K) P5 (atm) T5 (K) P5 (atm)1261 1.5 1488 1.51327 1.5 1548 1.51524 1.4 1607 1.5

(a) (b)

Figure 4.1 Measured (solid lines) and simulated (dashed lines) methanol and CO concentration time-

histories during the pyrolysis of methanol (time-zero: arrival of the reflected shock wave). Simulations used

the Princeton model (Li et al. [17]).

As shown in Figure 4.1(a), the simulated profiles from the model consistently

underpredict the methanol removal rates over the entire temperature range of the current

experiments, e.g., at 1 ms, the predicted methanol at 1458 K is 53% higher than that

measured. Similar underpredictions are observed in the CO concentration profiles shown

in Figure 4.1(b), which were measured during the pyrolysis of 0.2% methanol/Ar at

temperatures between 1403 and 1707 K and pressures between 1.1 and 1.4 atm. At 1 ms,

the Li et al. [17] mechanism underpredicted the CO concentration by 47% at 1507 K.

Separate experiments (for methanol and CO time-histories) were made using

different fuel concentrations in order to obtain high signal-to-noise ratio or to prevent

49

undesirably strong absorbance. In Figure 4.1, the methanol time-history data were

obtained using a 1% methanol/Ar mixture. Due to the endothermic nature of the

decomposition reaction, there is a small temperature drop in the reacting system,

perturbing the absorption cross-section coefficient methanol with time. By taking this

effect into account instead of assuming a constant methanol evaluated at the initial

temperature, more accurate species concentration time-histories are obtained. In this

study, the temperature profiles were calculated using the Li et al. [17] mechanism under

either constant enthalpy (H) and pressure (P) or constant energy (U) and volume (V). A

sample evaluation of methanol mole fraction time-history for 1% methanol/Ar mixture

initially at 1567 K and 2.1 atm was performed using three different approaches: constant

methanol (1567 K, 2.1 atm), T-dependent methanol based on constant (H, P) gasdynamic

model, and T-dependent methanol based on constant (U, V) gasdynamic model. As

demonstrated in Figure 4.2, the difference among these three calculated methanol

concentration time-histories is negligible; hereafter the constant methanol model, based on

the initial pressure and temperature behind the reflected shock is always employed for

methanol mole fraction conversion.

Among all contributing factors to the uncertainty in the experimental measurements,

the uncertainty in the post-shock temperature T5 most strongly affects the precision of the

measurements. Figure 4.3 illustrates the nominal predictions of methanol and CO

concentration profiles using the Li et al. [17] mechanism. The long-dashed lines

designate calculations at the nominal temperature of the experiments, whereas the short-

dashed lines represent those calculated with ±15 K uncertainty (1% uncertainty in T5).

The temperature uncertainty propagates to an uncertainty in species concentration

especially at the later times; i.e. at 1 ms, a 13% difference for methanol concentration and

20% for CO. It is also immediately clear from the comparisons in Figure 4.3 that the

model and experimental data cannot be reconciled through solely changing the simulated

T5 within its uncertainty bounds. Further discussion on uncertainty analysis of shock tube

datasets, which are shown to be a strong constraint on the kinetic model parameters, can

be found in [59].

50

(a) T5 =1567 K, P5 =2.1 atm, xmethanol =1% (b) T5 =1623 K, P5 =1.1 atm, xmethanol =0.2%

Figure 4.2 Comparison of (a) methanol and (b) CO concentration time-histories with different absorption

cross-sections in Beer’s law.

(a) (b)

Figure 4.3 Comparison of the measured (a) methanol and (b) CO time-histories with a detailed chemical

kinetic model. Long-dashed lines: predictions of the Li et al. [17] model at the nominal temperature shown;

short-dashed lines: computed uncertainty bounds due to ±15 K uncertainty in the T5 value.

Sensitivity analysis was first performed using the Li et al. [17] mechanism to

determine the reactions that are critical to controlling the species time-histories. The

sensitivity coefficient is the partial derivative of a species mole fraction with respect to

the reaction rate constant k, normalized by the maximum species mole fraction and the

parameter k. For instance, the CH3OH sensitivity SCH3OH is defined as:

maxCH3OH CH3OH CH3OHd d ,i iS t X X k k Eqn. 4-1

where XCH3OH is the local CH3OH mole fraction.

51

Figure 4.4 Sensitivity analysis (unmodified Li et al. [17] mechanism) of CH3OH at 100 s for 1% methanol

in argon at 1458 and 1567 K, respectively.

Figure 4.4 provides a CH3OH sensitivity plot at 100 s for a 1% methanol/Ar

mixture at 1458 K and 1567 K, respectively. At the early times, as expected, the

methanol concentration is most sensitive to the initial fuel decomposition:

CH3OH(+M) ↔ OH + CH3(+M), (1a)

However, an important methanol decomposition channel is not included in the Li et al.

[17] model:

CH3OH(+M) ↔ CH2(S) + H2O(+M). (1b)

This water elimination step results in the production of singlet methylene. GRI-Mech [60]

predicts this decomposition channel (1b) to account for ~40% of the initial methanol

consumption at 1200-1600 K, providing evidence that the reaction should be considered

in the current study of methanol decomposition. In fact, Jasper et al. [53] theoretically

calculated the rate constants for all the methanol decomposition channels and confirmed

reaction (1b) to be the second most important product channel. Hence, in the current

study, the Li et al. [17] methanol mechanism was modified to include two groups of

revisions. First, the rate constants for the methanol decomposition reaction (1a) in the Li

et al. [17] mechanism, and reaction (1b) that was added to the mechanism, were modified

using the values from Jasper et al. [53]. These modifications are summarized in Table 4.2.

Secondly, a series of singlet and triplet CH2 reactions required by the addition of reaction

(1b) was added; see Table 4.2. This set of reactions represents the minimal necessary set

of methylene reactions from the full GRI-Mech [60] methylene subset. That is, inclusion

52

of any additional methylene reactions from GRI-Mech [60] has negligible influence on

all modeling results presented in this work.

In terms of methanol unimolecular reactions (1a) and (1b), Jasper et al. [53]

estimated the 2 uncertainty factors to be 1.5-2 for their theoretical predictions of these

rates. In the current study, to achieve a best fit to the data, we found it necessary to

multiply the A-factor of the overall rate constant by a factor of 2, retaining their product

branching fractions. Given that the rates of these reactions are in the falloff regime for the

conditions of this work, we note that Master equation simulations in this region contain

additional uncertainty associated with collisional energy transfer. Although some

independently-corroborated experimental data exist for some conditions [61,62], there

still remains a large overall scatter in the data across all previous studies [53], and thus,

for this reason, modification of the overall rate constant of methanol decomposition

within a factor of 2 is justified. Figure 4.5 illustrates the influence of modified k1 on the

CH3OH and CO predictions by the Li et al. [17] mechanism. The modified mechanism is

able to accurately predict all the methanol time-histories especially at lower temperatures.

The early-time CO formation is also improved with k1 modified in the model.

0.0 0.5 1.0 1.5

0.0

0.5

1.0

1.5

1% Methanol/Ar

Measurement, current study Unmodified Li et al. mechanism Modified Li et al. (k1 modified only)

1610K, 2.2atm

1567K, 2.1atm

1368K, 2.4atm

1458K, 2.3atm

Time [ms]

Met

hano

l Mol

e Fr

actio

n [%

]

1266K, 2.5atm

(a)

0.0 0.5 1.0 1.5

0.00

0.05

0.10

0.15

0.20

0.25

0.2% Methanol/Ar

1403K, 1.2atm

1507K, 1.1atm

1707K, 1.3atm

Time [ms]

1623K, 1.1atm

CO

Mol

e Fr

actio

n [%

]

Measurement, current study Unmodified Li et al. mechanism Modified Li et al. (k1 modified only)

(b)

Figure 4.5 Influence of modified k1 (branching ratio from Jasper et al. [53]) on the (a) CH3OH and (b) CO

predictions by the Li et al. [17] mechanism. The spike at t = 0 is a result of beam steering from the detector

during the passage of the reflected shock and is not kinetic in nature.

53

Table 4.2 Reaction rate constants (near 1 atma) used in the current study: k = ATnexp(-Ea/RT)

Reaction A n Ea Fb Reference

CH3OH ReactionsCH3OH + M = CH3+OH+M k = k(P, T) 1.5-2 [53]c

CH3OH + M = CH2(S)+H2O+M k = k(P, T) 1.5-2 [53]c

CH3OH + H = CH2OH + H2 1.54E+06 2.35 5.91E+03 4 [63]CH3OH + H = CH3O + H2 5.48E+06 2.15 1.11E+04 4 [63]

CH2(S) ReactionsCH2(S)+Ar = CH2+Ar 2.41E+10 0.93 0.00 1.6 [60]CH2(S)+H2 = CH3+H 7.00E+13 0.00 0.00 3 [60]CH2(S)+H = CH+H2 3.00E+13 0.00 0.00 3 [60]CH2(S)+OH = CH2O+H 3.00E+13 0.00 0.00 3 [60]CH2(S)+CH3 = H+C2H4 1.20E+13 0.00 -5.70E+02 3 [60]CH2+CH2 = 2H+C2H2 2.00E+14 0.00 1.10E+04 3 [60]CH2+H2 = H+CH3 5.00E+05 2.00 7.23E+03 see text [60]CH2+CH3 = H+C2H4 4.00E+13 0.00 0.00 3 [60]CH2+H(+M) = CH3(+M) k∞

k0

6.00E+141.04E+26

0.002.76

0.001.60E+03

3 [60]

=0.562, T*** =9.10E+01, T* =5.83E+03, T** =8.55E+03CH2+OH = CH2O+H 2.00E+13 0.00 0.00 3 [60]CH2+OH = CH+H2O 1.13E+07 2.00 3.00E+03 3 [60]CH2+CH3OH = CH2OH+CH3 3.20E+01 3.2 7.17E+03 3 [64]CH2+CH3OH = CH3O+CH3 1.45E+01 3.1 6.94E+03 3 [64]CH+H2 = H+CH2 1.08E+14 0.00 3.11E+03 3 [60]CH+H2(+M) = CH3(+M) k∞

k0

1.97E+124.82E+25

0.43-2.80

-3.70E+025.90E+02

2 [60]

=0.578, T*** =1.22E+02, T* =2.53E+03, T** =9.36E+03CH+OH = H+HCO 3.00E+13 0.00 0.00 3 [60]CH+CH2 = H+C2H2 4.00E+13 0.00 0.00 6 [60]

CH3OCHO ReactionsCH3OCHO(+M)= CH3OH+CO(+M) 1.10E+13 0.00 5.87E+04 1.5-2 [65]d

CH3OCHO(+M) = CH4+CO2(+M) 7.50E+11 0.00 5.97E+04 see text [33]e

CH3OCHO(+M) = CH3O+HCO(+M) 8.36E+16 0.00 9.70E+04 see text [33]e

a T5 = 1260-1700 K. Rate constants are in s−1 and cm3 mole−1 s−1 for unimolecular and bimolecular reactions

respectively. Ea is in cal/mole.b Uncertainty factor.c Methanol decomposition is in the falloff region under the conditions of this work; see [53] for full k(T,P).

However, the A-factor has been increased by a factor of 2, with the branching ratio left unchanged.d Best fit to the experimental data in reference [65].e Only the high-pressure limit is shown.

54

Figure 4.6 Flux analysis of methanol pyrolysis at 800 s for 0.2% CH3OH/Ar at 1623 K and 1.1 atm.

However, as shown in Figure 4.5(b), the modified Li et al. [17] mechanism (k1

modified only) still underpredicts the CO concentration time-histories, especially at

higher temperatures. In addition, the final CO plateau (see the case at 1623 K) is about 5%

lower than the measurement and the unmodified Li et al. [17] mechanism. Flux analysis

shown in Figure 4.6 reveals that most methanol (>80%) is consumed through H-

abstraction reactions by atomic hydrogen:

CH3OH + H ↔ CH2OH + H2, (2a)

↔ CH3O + H2, (2b)

where k2a is the dominant reaction. The oxygen atoms in reaction (2a) and (2b) finally

form CO primarily through subsequent H-abstraction reactions. With methanol

unimolecular decomposition rate constants modified in the Li et al. [17] mechanism,

more oxygen atoms go on to form H2O since channel (1b) forms H2O directly and

channel (1a) produces the OH radical, which subsequently consumes methanol to

produce H2O through H-abstraction reactions. Therefore, we conclude that the H-

abstraction reactions of methanol by atomic hydrogen in the Li et al. [17] mechanism

may need revision to achieve accurate CO production.

In the Li et al. [17] mechanism, the overall rate constant k2 used the value

recommended by Warnatz [66], which is a factor of 2.6-5 times lower than that in GRI-

Mech [60] over the temperature range of 1200-1800 K. Meana-Pañeda et al. [63] recently

performed direct-dynamics variational transition state theory calculations of CH3OH+H

54

Figure 4.6 Flux analysis of methanol pyrolysis at 800 s for 0.2% CH3OH/Ar at 1623 K and 1.1 atm.

However, as shown in Figure 4.5(b), the modified Li et al. [17] mechanism (k1

modified only) still underpredicts the CO concentration time-histories, especially at

higher temperatures. In addition, the final CO plateau (see the case at 1623 K) is about 5%

lower than the measurement and the unmodified Li et al. [17] mechanism. Flux analysis

shown in Figure 4.6 reveals that most methanol (>80%) is consumed through H-

abstraction reactions by atomic hydrogen:

CH3OH + H ↔ CH2OH + H2, (2a)

↔ CH3O + H2, (2b)

where k2a is the dominant reaction. The oxygen atoms in reaction (2a) and (2b) finally

form CO primarily through subsequent H-abstraction reactions. With methanol

unimolecular decomposition rate constants modified in the Li et al. [17] mechanism,

more oxygen atoms go on to form H2O since channel (1b) forms H2O directly and

channel (1a) produces the OH radical, which subsequently consumes methanol to

produce H2O through H-abstraction reactions. Therefore, we conclude that the H-

abstraction reactions of methanol by atomic hydrogen in the Li et al. [17] mechanism

may need revision to achieve accurate CO production.

In the Li et al. [17] mechanism, the overall rate constant k2 used the value

recommended by Warnatz [66], which is a factor of 2.6-5 times lower than that in GRI-

Mech [60] over the temperature range of 1200-1800 K. Meana-Pañeda et al. [63] recently

performed direct-dynamics variational transition state theory calculations of CH3OH+H

54

Figure 4.6 Flux analysis of methanol pyrolysis at 800 s for 0.2% CH3OH/Ar at 1623 K and 1.1 atm.

However, as shown in Figure 4.5(b), the modified Li et al. [17] mechanism (k1

modified only) still underpredicts the CO concentration time-histories, especially at

higher temperatures. In addition, the final CO plateau (see the case at 1623 K) is about 5%

lower than the measurement and the unmodified Li et al. [17] mechanism. Flux analysis

shown in Figure 4.6 reveals that most methanol (>80%) is consumed through H-

abstraction reactions by atomic hydrogen:

CH3OH + H ↔ CH2OH + H2, (2a)

↔ CH3O + H2, (2b)

where k2a is the dominant reaction. The oxygen atoms in reaction (2a) and (2b) finally

form CO primarily through subsequent H-abstraction reactions. With methanol

unimolecular decomposition rate constants modified in the Li et al. [17] mechanism,

more oxygen atoms go on to form H2O since channel (1b) forms H2O directly and

channel (1a) produces the OH radical, which subsequently consumes methanol to

produce H2O through H-abstraction reactions. Therefore, we conclude that the H-

abstraction reactions of methanol by atomic hydrogen in the Li et al. [17] mechanism

may need revision to achieve accurate CO production.

In the Li et al. [17] mechanism, the overall rate constant k2 used the value

recommended by Warnatz [66], which is a factor of 2.6-5 times lower than that in GRI-

Mech [60] over the temperature range of 1200-1800 K. Meana-Pañeda et al. [63] recently

performed direct-dynamics variational transition state theory calculations of CH3OH+H

55

reactions using the microcanonically optimized multidimensional tunneling transmission

coefficient, showing good agreement with the theoretical study by Jodkowski et al. [67].

These studies of the rate constant k2 and the corresponding branching ratio are plotted in

Figure 4.7 for comparison. At temperatures between 1200 and 1800 K, a temperature-

dependent branching ratio (0.84-0.91) is recommended by Meana-Pañeda et al. [63]

instead of the constant value of 0.8 used in the Li et al. [17] mechanism and GRI-Mech

[60]. Therefore, we modified k2 in the Li et al. [17] mechanism using the values

calculated by Meana-Pañeda et al. [63], with recognition that the uncertainty for this

reaction remains high (F = 4 at 2500 K for both 2a and 2b) because neither uncertainty

analysis nor estimation was performed by the authors. Furthermore, there remains a large

spread between many rate determinations under the temperature range of this study.

Specifically, the discrepancies between the data of Cribb et al 1992 [68] and Vandooren

et al. 1981 [69] were not critically reviewed or discussed. Thus, the assigned uncertainty

here reflects that recommended by Baulch at 2500 K [70]. The uncertainty in the

branching ratio, however, is smaller than the absolute overall rate constant and likely less

than 15% of the temperature-dependent values computed by Meana-Paneda et al. [63].

Figure 4.7 Reaction rate constants of CH3OH+H (k2) and branching ratio.

The complete revisions to the Li et al. [17] mechanism, including the revisions

associated with the methanol decomposition reactions (k1) and the H-abstraction reactions

(k2), are shown in Table 4.2. Figure 4.8 compares the measured CH3OH and CO time-

56

histories with the new simulated results. Improved agreement is seen for both CH3OH

and CO profiles over the entire temperature range.

Small discrepancies in the CO profiles remain. Hence, further study of the reactions

involving hydroxymethyl and formaldehyde may be necessary to correct this problem.

For instance, the early-time CO formation is also sensitive to hydroxymethyl reactions

(CH2OH ↔ CH2O + H, CH2OH +H ↔ CH3 + OH) and the subsequent H-abstraction

from formaldehyde (CH2O + H ↔ HCO + H2), as shown in the CO sensitivity plot of

Figure 4.9. In the modified Li et al. [17] model, we leave these rate constants unchanged.

However, very different rate constant values are assigned to these reactions in different

models. For instance, the Li et al. [17] mechanism used the recommended k4 from the

pyrolysis study of methanol by Cribb [68], which is a factor of 10 times lower compared

to the value in GRI-Mech [68]. Accurate electronic structure and master equation

calculations for this reaction are recommended to achieve further reduction of

uncertainties in the C1 combustion mechanism.

0.0 0.5 1.0 1.5

0.0

0.5

1.0

1.5

1610K, 2.2atm

Measurement, current study Modified Li et al. (k1 and k2 modified)

1567K, 2.1atm

1368K, 2.4atm

1458K, 2.3atm

Time [ms]

Met

hano

l Mol

e Fr

actio

n [%

]

1266K, 2.5atm

1% Methanol/Ar

(a)

0.0 0.5 1.0 1.5

0.0

0.5

1.0

1.5

2.0

2.5

1403K, 1.2atm

1507K, 1.1atm

1707K, 1.3atm

Time [ms]

1623K, 1.1atm

CO

Mol

e Fr

actio

n [%

]

Measurement, current study Modified Li et al. (k1 and k2 modified)

0.2% Methanol/Ar

(b)

Figure 4.8 Effect of modifications to the Li et al. [17] mechanism predictions for (a) CH3OH and (b) CO

concentration time-histories during the pyrolysis of methanol.

57

Figure 4.9 Sensitivity analysis (Li et al. [17] mechanism with k1 modified) of CO concentration at 100 s

for 0.2% CH3OH/Ar at 1507 K and 1623 K.

4.4 High-Temperature Methyl Formate Pyrolysis

As discussed before, the chemical kinetics of methyl formate has been studied by

several groups both theoretically and experimentally [33,54,71–74]. Dooley et

al. [33] constructed a detailed kinetic model for methyl formate combustion, which was

tested against the experimental data obtained in three different systems: a turbulent flow

reactor, a shock tube, and a laminar MF/air flame. This mechanism has recently been

used to simulate a low-pressure (22–30 torr) burner-stabilized laminar flame [71],

showing general agreement of model versus experiment. However, according to the

authors, no definite conclusion on the kinetics of methyl formate decomposition was

reached considering the difference of the rate constants used in the model [33] and the

recent theoretical calculations [73], especially at lower pressures. The dominant

decomposition channel of methyl formate has been accepted in general to be:

CH3OCHO(+M) ↔ CH3OH + CO(+M), (3a)

along with two other molecular reactions, which are minor channels in MF

decomposition:

CH3OCHO(+M) ↔ CH4 + CO2(+M), (3b)

CH3OCHO(+M) ↔ 2CH2O(+M). (3c)

Besides these molecular channels, two radical-related decomposition pathways were also

included in the Dooley et al. [33] mechanism:

58

CH3OCHO(+M) ↔ CH3O + HCO(+M), (3d)

CH3OCHO(+M) ↔ CH3 + OCHO(+M). (3e)

with rate constants orders of magnitude smaller than k3a.

Example CO concentration time-histories measured using mid-IR absorption near

4.56 m are shown in Figure 4.10. Interfering absorption due to other hydrocarbon

intermediates in the early decomposition is negligible. In comparing these experimental

data and others with predictions of the Dooley et al. [33] mechanism (see Figure 4.10),

the calculations were carried out for homogeneous, adiabatic conditions with a constant-

internal-energy, constant-volume constraint using the CHEMKIN-PRO [34] software

package. Discrepancies between model and experiment, beyond experimental uncertainty,

are evident at long times and especially at higher temperatures. In addition, Dooley’s

mechanism [33] was revised with k3a-k3c modified using the calculations by Metcalfe et al.

[73] and the simulated results are also illustrated in Figure 4.10 for comparison. However,

the revised mechanism poorly predicted the experimental results over the entire

temperature range. Therefore, all of our kinetic analysis is based on the Dooley et al. [33]

mechanism.

Figure 4.10 Representative CO concentration time-histories measured during the decomposition of MF at

various temperatures under a fixed initial fuel concentration (0.1% MF/Ar) compared with the predictions

of the Dooley et al. [33] mechanism and that with k3a-k3c revised from Metcalfe et al. [73].

59

Figure 4.11 Local sensitivity analysis for CO concentration using the Dooley et al. [33] mechanism (0.1%

MF/Ar, 1376 K, 1.58 atm).

Sensitivity analysis was performed to identify the dominant reactions that affect

species time-histories in the system. As expected, the sensitivity analysis shown in Figure

4.11 confirms that the CO concentration time-history is predominantly controlled by k3a

(Rxn. 1 in Figure 4.11). By modifying only k3a in the chemical kinetic model, we obtain

excellent fits to the experimental CO time-histories as demonstrated in Figure 4.12. The

best-fit k3a for the example case is determined to be 5.3×103 s-1 (compared to 4.6×103 s-1

in the Dooley et al. [33] mechanism, a difference of 15%) with an estimated fitting error

of less than ±10%. Notice that the rate constant was fit only to the early-time (100 s in

Figure 4.12) behavior of the CO trace; at later times, some sensitivity to secondary

reactions exists.

The rate constant determinations for reaction (3a) are illustrated in Figure 4.13 on an

Arrhenius plot along with a least-squares fit. The best-fit to the current data yields a first-

order rate constant expression for k3a valid over the range 1187-1607 K, 1.46-1.72 atm:

k3a = 1.1×1013 exp(-29560/T, K) s-1, where the root-mean-square (RMS) experimental

scatter about the fit is 5.5%. The primary contributions to uncertainties in the rate

constants are: temperature (10%), fitting the data to computed profiles (5%), and CO

cross-section (3%). These uncertainties give conservative overall uncertainties in k3a of

35%. The current experimental results fall between previous rate values from Westbrook

60

et al. [75], Dooley et al. [33], Peukert et al. [74] and Metcalfe et al. [73], and are in good

agreement with the rate constants used in the Dooley et al. [33] mechanism.

Figure 4.12 Example MF decomposition k3a rate constant determination. Solid black line, experimental data;

solid red line, best fit to the data with the optimal value of k3a; dashed lines, variation of k3a±50%.

Figure 4.13 Comparison of measured k3a (1.5-1.7 atm) with previous rate constants (LLNL [75], Princeton

[33], Argonne [74] and NUI [73]) for the reaction CH3OCHO→ CH3OH + CO. Least-squares fit (in black)

to experimental data gives k3a = 1.11013 exp(-29556/T, K) s-1.

Pressure dependence of (Rxn. 3a) was also investigated by measuring the rate

constants at lower pressures (0.3-0.4 atm) and higher pressures (4.6-5.2 atm) as shown in

Figure 4.14. All the experimental results are also tabulated in Table 4.3. A weak pressure-

dependence of k3a was observed in the experiments for pressures higher than 2 atm. This

plot supports our expectation that measurements in the current study (~1.6 atm) are

61

relatively close to the high-pressure limit, and hence first-order coefficients are reported.

Again, good agreement can be found between our measurements and k3a used in the

Dooley et al. [33] mechanism. Reaction (3a) has also been theoretically studied recently

by Peukert et al. [74] at the theory level of CCSD(T)/cc-PV∞Z. The Master equation

calculations of rate constants (1000-2000 K) were performed by considering 1-D

hindered rotor treatments, tunneling corrections and a temperature-dependent energy

transfer parameter <E>down. Although their calculated values of k3a∞ are in good

agreement (±30%) with the k3a∞ recommendations of Metcalfe et al. [73], the values of

k3a(1 atm) in Peukert et al. [74] are 1.5-5 times the predictions of k3a(1 atm) predictions of

Metcalfe et al. [73] over 1000-2000 K. The calculations by Peukert et al. [74] are in good

agreement with our measurements (within 50%) at pressures near 5 atm, but varied by a

factor of 2 or more at pressures lower than 1 atm. Further study to reduce the uncertainty

for this rate constant is still in progress. Recent preliminary calculations from Argonne

National Lab show excellent agreement with our results by adjusting the barrier height in

[74] by ~0.5 kcal/mol.

Figure 4.14 Summary of recent studies of k3a. Symbol: shock tube measurement; dashed line: Peukert et al.

[74]; dash-dot line: Metcalfe et al. [73].

62

Table 4.3 Test conditions and rate constant data for reaction: CH3OCHO→ CH3OH + CO.

T (K) P (atm) k3a (s-1) T (K) P (atm) k3a (s-1) T (K) P (atm) k3a (s-1)1189 0.43 1.3102 1202 1.72 2.1102 1242 5.26 7.5102

1215 0.40 2.1102 1241 1.67 4.8102 1277 5.05 1.3103

1226 0.38 2.7102 1285 1.65 1.2103 1354 4.82 4.8103

1241 0.38 3.9102 1309 1.62 1.8103 1450 4.60 2.0104

1279 0.36 6.2102 1340 1.58 2.9103 1539 4.44 6.5104

1303 0.33 9.5102 1342 1.61 3.1103

1340 0.32 1.6103 1354 1.62 3.9103

1354 0.32 2.1103 1376 1.58 5.3103

1395 0.30 3.8103 1428 1.54 1.1104

1440 0.43 9.1103 1478 1.49 2.2104

1459 0.41 1.2104 1488 1.51 2.6104

1534 0.38 2.5104 1512 1.53 3.3104

1602 0.33 4.8104 1548 1.50 5.4104

1608 0.33 5.1104 1556 1.46 6.5104

1607 1.47 1.2105

Methanol is a major intermediate during MF thermal decomposition as it is also

produced from reaction (3a). Hence the availability of a methanol diagnostic provides

another validation opportunity of the MF mechanism. Furthermore, because the Dooley

et al. [33] mechanism adopts the C1 combustion model by Li et al. [17], the effects of the

modified methanol mechanism (discussed in Section 4.3) on predictions of species time-

histories during MF pyrolysis can be examined.

The measured methanol concentration time-histories during the pyrolysis of 1%

methyl formate in argon are plotted in Figure 4.15, along with the model predictions from

the original and the modified Dooley et al. [33] mechanisms. Note that only the rate

constants for methanol decomposition (k1 and k2) and MF unimolecular elimination (k3a)

were modified as discussed before. At temperatures between 1261 K and 1524 K, the

unmodified Dooley et al. [33] model is in good agreement with the current measurements

regarding the early-time methanol formation, which is determined by (Rxn. 3a). However,

the model fails to predict the removal rate of methanol at later times; see the case at 1524

K after 0.1 ms. In contrast, the modified mechanism accurately predicts the full methanol

time-histories (both formation and removal rates) at all the temperatures studied.

63

(a) Original Dooley et al. mechanism

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

Met

hano

l Mol

e Fr

actio

n [%

]

1261K, 1.5atm

1327K, 1.5atm

1524K, 1.4atm

Time [ms]

Measurement, current study Modified Dooley et al. (k1+k2 modified)

1% MF/Ar

(b) Modified Dooley et al. mechanism

Figure 4.15 Comparisons of measured and simulated methanol time-histories for 1% methyl formate in

argon. Only the reaction rate constants k1, k2 and k3a are modified in the Dooley et al. [33] mechanism.

Figure 4.16 Example CO concentration time-histories: solid line, measurement; dashed line, simulation

using unmodified Dooley et al. [33] mechanism; dash-dot line, simulation using the Dooley et al.

mechanism with k1, k2 and k3a modified.

Similarly, an obvious discrepancy still exists at long times of CO time-histories

despite the good agreement in the early-time CO formation. A sample CO profile during

the pyrolysis of 0.1% MF/Ar at 1607 K and 1.5 atm is shown in Figure 4.16. Two-stage

CO formation is observed during the thermal decomposition of methyl formate. The rapid

early-time (prior to approximately 50 s) CO formation is directly from the initial fuel

decomposition (Rxn. 3a), which is well captured by the Dooley et al. [33] mechanism. At

long times, however, the simulated CO profile from the Dooley et al. [33] mechanism is

inconsistent with the measurement. At 1 ms, for instance, the model significantly

64

underpredicts the CO plateau level by 15%. On the other hand, the simulated CO time-

history with k1 and k2 modified in the model is in better agreement with the measurement.

However, at times between 50 and 400 s, there still exists a gap between the measured

and simulated CO concentration time-histories that is somewhat greater than the

experimental uncertainty in the CO data.

Figure 4.17 CO sensitivity (Dooley et al. [33] mechanism) for 0.1% MF/Ar, 1607 K and 1.5 atm.

Figure 4.18 Reaction rate constants k3a, k3b and k3d in the Dooley et al. [33] mechanism.

Finally, it is of interest to briefly discuss the underlying chemistry that is responsible

for the remaining discrepancy between the simulated CO time-histories using the

modified Dooley et al. [33] mechanism and the measurements. Figure 4.17 provides a

sensitivity plot of CO concentration calculated at the same condition as shown in Figure

4.16. Of the MF decomposition reactions, the early-time CO formation is dominantly

controlled by (Rxn. 3a) to produce CH3OH+CO (channel F in Figure 4.17). However,

65

after about 50 s, CO sensitivity is complicated by two other MF decomposition

reactions (channels G and H in Figure 4.17) and multiple secondary reactions. However,

those secondary reactions are directly related to the methanol sub-mechanism as

discussed in Section 4.3. With k1 and k2 modified, the model is able to predict the CH3OH

and CO time-histories reasonably well during the methanol thermal decomposition.

Hence, the remaining uncertainty in the Dooley et al. [33] mechanism is considered to be

mainly from the MF decomposition channels:

CH3OCHO(+M) ↔ CH4 + CO2(+M), (3b)

CH3OCHO(+M) ↔ CH3O + HCO(+M), (3d)

for which the rate constants are orders of magnitude smaller than (Rxn. 3a) in the Dooley

et al. [33] mechanism, as shown in Figure 4.18.

The molecular channel (3b) produces CH4+CO2 through a four-membered ring

transition state, competing with channel (3a) to form CO. This channel plays a more

important role at the later times of CO time-histories as a negative factor. In contrast, the

radical-related channel (3d) directly creates CH3O and HCO through bond scissions,

which then quickly form CO through pathways CH3O→CH2O→HCO→CO. Thus, rate

adjustments of these two reactions do not significantly affect the initial fuel

decomposition rate, but will change the CO formation at the later times. Here with the A-

factor of k3b halved and that of k3d increased by a factor of two in the modified Dooley et

al. [33] mechanism, the simulated CO profiles at three different temperatures (1488 K,

1548 K, and 1607 K) are illustrated in Figure 4.19, showing good agreement with the

measurements over the entire temperature range.

Due to how these rate constants were determined, their uncertainties are certainly a

factor of 2 or more. As described by Dooley et al. in their work [33], reaction (3d) was

determined by estimating the high-pressure limit rate of the reverse reaction and using the

Sumathi and Green [76] derived thermochemistry to determine the decomposition step.

The high-pressure limit rate for (3b) was estimated based on a number of assumptions

and analogies. Furthermore, lack of details regarding the ME/QRRK calculations make it

difficult to quantify the uncertainties resulting from the low pressure limit and Troe

parameters. Lastly, due to the lack of direct experimental evidence and the complex

66

nature of MF decomposition, the uncertainties for rate constants k3b and k3d as determined

in Dooley et al. [33] remain high, and the adjustments here are justified.

Since the simulated methanol profiles using unmodified MF decomposition rate

constants are in good agreement with the current measurement (see Figure 4.15, right

panel), the issue arises as to whether the adjustment of k3b and k3d may change the

methanol consumption rate, as the atomic hydrogen formed through the subsequent

decomposition of CH3O and HCO accelerates the H-abstraction reaction of methanol.

The simulated methanol time-histories using the modified Dooley et al. [33] mechanism

with and without the MF decomposition rate constants (k3) modified are illustrated in

Figure 4.20. The difference of the methanol time-histories between these two modified

models is almost negligible. Therefore, the modified Dooley et al. [33] mechanism with

k1, k2 and k3 modified can accurately predict both methanol and CO concentration time-

histories throughout the current experimental conditions studied.

Figure 4.19 Effect of modifications to the Dooley et al. [33] model predictions for the CO concentration

time-histories during the pyrolysis of methyl formate.

Figure 4.20 Comparisons of measured and simulated methanol time-histories during MF pyrolysis.

0.0 0.2 0.4 0.6 0.8 1.00.00

0.05

0.10

0.15

0.20

0.25

1488K, 1.5atm

1548K, 1.5atm1607K, 1.5atm

CO

Mol

e Fr

actio

n [%

]

Time [ms]

0.1% MF/Ar

Measurement by Ren et al. 2012 Modified Dooley et al. (k1+k2+k3 modified)

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

1.2

Met

hano

l Mol

e Fr

actio

n [%

]

Time [ms]

Measurement, current study Unmodified Dooley et al. 2010 Modified Dooley et al. (k1+k2 modified) Modified Dooley et al. (k1+k2+k3 modified)

1% MF/Ar1524 K, 1.4 atm

67

Chapter 5. Thermal Decomposition of C3-C5

Ethyl Esters

5.1 Introduction

As discussed previously, biodiesel is typically produced through the

transesterification of vegetable oils or animal fats with methanol yielding fatty acid

methyl esters (FAMEs) [77,78]. These methyl esters can be blended with petroleum

diesel and used in diesel engines without major modifications. Methanol instead of

ethanol has been preferred in industrial application to produce biodiesel mostly due to its

lower cost. However, ethanol is less toxic, less volatile and less corrosive than methanol,

and therefore provides a safer work environment during the transesterification process.

Additionally, some countries like Brazil have started to produce ethanol in large

quantities [79]. Hence, biodiesel in the form of fatty acid ethyl esters (FAEEs) produced

through the conversion of biolipids with ethanol would further enhance the sustainability

and commercialization of biofuels [80].

Previous kinetic studies on ethyl esters have aimed to compare the small methyl and

ethyl esters with the same chemical formula (isomers) while varying the length of the

alkyl chains to investigate the effect of the molecular structure on the combustion

chemistry. In 2009, Westbrook et al. [75] developed a detailed chemical kinetic

mechanism describing the laminar premixed flames of four small alkyl ester fuels: methyl

formate, methyl acetate, ethyl formate and ethyl acetate. The model development

employed a principle of similarity of functional groups in constraining the H-atom

abstraction and unimolecular decomposition reactions for each of these esters. Akih-

Kumgeh and Bergthorson [81] investigated the ignition behavior of three pairs of

68

methyl/ethyl esters, including methyl/ethyl formate, methyl/ethyl acetate, and

methyl/ethyl propanoate, by measuring the ignition delay times behind reflected shock

waves. Ethyl esters are generally characterized by shorter ignition delay times than those

of methyl esters. Metcalfe et al. [82] performed an experimental and modeling study of

C5H10O2 ethyl and methyl ester oxidation. The detailed kinetic model describing ethyl

propanoate and methyl butanoate oxidation was validated against the ignition delay times

for a series of mixtures ( = 0.25-1.5) behind reflected shock waves (1100-1670 K, 1.0

and 4.0 atm). This work was complemented later by jet-stirred reactor (JSR) experiments

and modeling by Metcalfe et al. [83]. The study by Walton et al. [84] for methyl

butanoate and ethyl propanoate combustion refined the mechanism of Metcalfe et al. [82],

so that it could reproduce new experimental results from a rapid compression machine. In

2011, Yang et al. [85] performed a low-pressure flame study of three C5H10O2 ester

flames: methyl butanoate, methyl isobutanoate, and ethyl propanoate. A detailed kinetic

mechanism was constructed to describe differences in the compositions of key reaction

intermediates between the flames of these ester isomers. Very recently, Dayma et al. [86]

investigated the laminar burning velocities of C4-C7 ethyl esters at 1, 3, 5 and 10 bar in a

spherical combustion chamber.

The main purpose of this work is to provide new experimental results of the thermal

decomposition of ethyl esters behind reflected shock waves and their kinetic

interpretation. We have measured CO, CO2 and H2O concentration time-histories using

laser-absorption techniques during the pyrolysis of three ethyl esters: ethyl formate (EF,

C3H6O2), ethyl acetate (EA, C4H8O2), and ethyl propanoate (EP, C5H10O2). Figure 5.1

shows their corresponding molecular structures. Considering the different groups (-H, -

CH3 and -CH2CH3) adjacent to C=O function, our experimental results should reveal the

effect of varying the alkyl chain length on the decomposition characteristics of small

ethyl esters.

(a) C3H6O2 (b) C4H8O2 (c) C5H10O2

Figure 5.1 The molecular structures of (a) ethyl formate (b) ethyl acetate and (c) ethyl propanoate.

69

5.2 Experimental

5.2.1 Shock Tube and Laser Diagnostics

All the experiments were performed in the same shock tube (15.24 cm inner

diameter) as introduced in Chapter 2 and Chapter 4. All fuels (>99% pure, Sigma-Aldrich)

were frozen and degassed three times to remove dissolved volatiles before making the

mixtures. All the test mixtures were manometrically prepared in a stainless-steel mixing

tank (40 L) heated uniformly to 50°C with an internal magnetically driven stirrer. Laser

absorption and side-wall pressure measurements (Kistler 601B1 PZT) were located 2 cm

from the shock tube end wall.

In this study, three laser absorption diagnostics were utilized for the time-resolved

measurements of CO, H2O and CO2 concentration time-histories. CO concentration was

measured using a DFB-QC laser, detecting the peak intensity of the CO R(13) absorption

line at 2193.36 cm-1. Absorption measurements of H2O were performed using a

distributed feedback (DFB) diode laser at 2550.96 nm within the ν3-fundamental

vibrational band, achieving a minimum H2O detection sensitivity of 25 ppm at 1400 K

and 1.5 atm for a path length of 15 cm [37].

A new mid-IR CO2 diagnostic was developed in this work by incorporating an

external cavity quantum cascade laser (ECQCL), to provide sensitive and quantitative

measurements of carbon dioxide. The R(76) transition line in the CO2 fundamental band

near 4.3 m was selected due to its high absorption strength and negligible interference

from other combustion products. The Ar-broadening coefficient (2CO2-Ar) for this

transition was measured behind reflected shock waves using the same method as

described in reference [37]. The Ar-broadening coefficient over the temperature range of

1200-1900 K was measured to be 0.0762±0.0012 cm-1/atm with a temperature exponent

of n=0.57±0.02. Compared to previous CO2 sensors detecting the overtone and

combinational bands near 2.7 m [40,87], this new diagnostic scheme provides orders-of-

magnitude greater sensitivity for shock tube experiments.

70

5.2.2 Experimental Results

The exact experimental conditions behind reflected shock waves are summarized in

Table 5.1. The measurements covered the temperature range of 1301-1636 K and

pressure range of 1.48-1.72 atm with fuel concentration of 2000 ppm. Such dilute

mixtures result in negligible temperature variation during fuel pyrolysis. Hence, no

correction is needed for laser absorption coefficients.

The representative H2O, CO and CO2 concentration time-histories during EF, EA

and EP pyrolysis are presented in Figure 5.2 at temperatures near 1450 K and pressures

near 1.5 atm. In general, the species time-histories for these ethyl esters differ

significantly from each other. EF shows the highest yield of CO and H2O (with nearly

equal formation rate) among these three esters, but produces a negligible amount of CO2.

The CO profile during EA pyrolysis exhibits the slowest formation rate among all these

esters. Additionally, CO, H2O and CO2 are observed to have almost the same formation

rate during the pyrolysis of EP.

The product fractional yields (defined by xproduct/xfuel, here xfuel = 2000 ppm) at t = 1

ms for the C3-C5 ethyl esters are plotted as a function of temperature in Figure 5.3. It is

obvious that these esters demonstrate completely distinct product yield behaviors, but the

final O-atom carrying products are mainly found to be CO, H2O and CO2. At the highest

temperature when the product time-histories reached the plateau level within the test time

of 2 ms, the O-atom balance in the measured CO, H2O, and CO2 profiles were counted to

be 98%, 93% and 95% for EF, EA and EP, respectively. Note that all the detailed species

time-history data will be presented and discussed in Section 5.4.

(a) (b) (c)

Figure 5.2 Measured species time-histories during the pyrolysis of (a) EF (b) EA and (c) EP at temperature

near 1450 K and pressures near 1.5 atm, with fuel concentration 2000 ppm in argon.

71

(a) Ethyl formate (b) Ethyl acetate (c) Ethyl propanoate

Figure 5.3 Measured product fractional yield for (a) EF (b) EA and (c) EP at t = 1 ms.

Table 5.1 Summary of reflected shock conditions for ethyl ester pyrolysis.

5.3 Kinetic Modeling

Kinetic modeling of the shock tube species time-histories is carried out to gain

further insight into the pyrolysis of EF, EA and EP using the CHEMKIN code [34]. A

detailed chemical kinetic mechanism (139 species and 786 reactions) for EP oxidation

has been developed by Metcalfe et al. [83]. This EP oxidation model was able to simulate

shock tube ignition delay times (1100-1670 K, 1 and 4 atm [82]) and jet-stirred reactor

data (750-1100 K, 10 atm [83]). Under the pyrolysis conditions, the major EP destruction

pathways using the Metcalfe et al. [83] mechanism are summarized in Figure 5.6. Nearly

all EP decomposes to ethylene and propanoic acid through the concerted dissociation

Mixture CO diagnostic H2O diagnostic CO2 diagnostic

2000ppm

EF/Ar

T5 (K) P5 (atm) T5 (K) P5 (atm) T5 (K) P5 (atm)

1314 1.69 1314 1.69 1370 1.65

1402 1.56 1402 1.56 1449 1.63

1467 1.68 1467 1.68 1636 1.49

1629 1.48 1629 1.48

2000ppm

EA/Ar

1472 1.48 1393 1.54 1492 1.61

1566 1.52 1472 1.48 1578 1.54

1634 1.49 1566 1.52 1634 1.49

2000ppm

EP/Ar

1310 1.58 1310 1.58 1301 1.72

1351 1.56 1351 1.56 1366 1.69

1440 1.56 1440 1.56 1454 1.57

1567 1.51 1567 1.51 1580 1.52

72

pathway (EP = C2H4 + C2H5COOH). The produced propanoic acid is then consumed

mainly via the H-atom abstraction reactions yielding CH2CH2COOH and CH3CHCOOH

radicals. Subsequent -scissions of these two radicals produce stable intermediate and

smaller radicals such as ethylene, methyl ketene, hydroxyl radical, and HOCO radical.

Yang et al. [85] recently studied the low-pressure premixed flat flames of ethyl

propanoate, reporting that the rate constants of EP six-center unimolecular elimination

and H-atom abstraction reactions needed to be modified for better agreement with the

measured composition of reaction intermediates in the low-pressure flames. In the current

study, the Metcalfe et al. [83] mechanism combined with the recommended modifications

by Yang et al. [85] is used for analyzing the experimental results of EP pyrolysis.

There are no mechanisms optimized for EF and EA pyrolysis, so the kinetic

modeling in this work includes the construction of EF and EA pyrolysis models in a

hierarchical manner. The core C1-C4 kinetic submechanisms from Metcalfe et al. [83]

are used as the starting point for the current EF and EA mechanisms. Here we only

discuss the modeling efforts in terms of the added reactions or modified reaction rate

constants in the Metcalfe et al. [83] mechanism. The thermodynamic data for radicals

related to EF and EA were taken from the literature and compared with the values

estimated using THERM [88] and THERGAS [89] codes.

Figure 5.4 presents the major EF destruction pathways. The initial EF decomposition

is generally accepted to be the unimolecular elimination (EF = C2H4 + HCOOH) through

a six-center transition state [90], followed by formic acid decomposition taking two

competing pathways of dehydration (HCOOH = H2O + CO) and decarboxylation

(HCOOH = CO2 + H2). The EF submechanism is taken directly from Westbrook et al.

[75], including the unimolecular decomposition to stable molecules or radicals through

bond cleavage, and the H-atom abstractions by H, OH and CH3 radicals. As discussed

before, the alcohol elimination reaction is the dominant pathway during the pyrolysis of

methyl formate [65,72,73]. Considering the similar molecular structure between methyl

and ethyl formate, hence, another concerted unimolecular reaction (EF = CO + C2H5OH)

is added into the EF submechanism with the rate constant estimated by analogy with

methyl formate [65]. The branching ratio of this ethanol elimination reaction is evaluated

73

to be 0.03 and 0.1 at 1200 K and 1800 K, respectively, proving to be a minor channel

during EF pyrolysis.

Formic acid is the major intermediate during EF pyrolysis, and its decomposition

has been the subject of several research groups [91–93]. Our current shock tube

measurements result in CO2/H2O ratios between 0.05 and 0.07, which is in excellent

agreement with the experimental observation reported by Saito et al. [93]. Hence, the rate

expressions recommended by Saito et al. [93] were used in the current EF submechanism.

The H-atom abstraction reactions are also the possible consumption pathways for formic

acid, with rate constants taken from [94]. The entire EF submechanism is summarized in

Table 5.2.

In the case of EA pyrolysis, the unimolecular decomposition of EA produces one

ethylene molecule and one acetic acid molecule via a six-center transition state, followed

by the subsequent decomposition of acetic acid to the final products of H2O, CO, CO2

and methane. The major destruction pathways for EA are described in Figure 5.5.

Similarly, the EA submechanism is taken directly from Westbrook et al. [75], including

EA unimolecular decomposition reactions and the H-atom abstraction by H, OH and CH3

radicals. Acetic acid is the major intermediate species during EA pyrolysis. Leplat and

Vandooren [95] recently performed numerical and experimental study of the combustion

of acetic acid in three CH3COOH/O2/Ar flat premixed flames burning at low pressure (50

mbar) and with equivalence ratios equal to 0.77, 0.9 and 1.05, respectively. Therefore, in

this work, the CH3COOH submechanism from Leplat and Vandooren [95] has been

added to the current EA pyrolysis model; see Table 5.3 for all the added reactions.

C2H4 +

H2O + CO

CO2 + H2

H

HO

CO

H

O

O HO

O

H

CH2

CH2

Figure 5.4 EF pyrolysis: major destruction pathways.

74

C2H4 +

H2O +

CO2 + CH4

CO

Figure 5.5 EA pyrolysis: major destruction pathways.

+OH

C2H4 +HOCO + C2H4-H

-H

Figure 5.6 EP pyrolysis: major destruction pathways.

5.4 Discussion

5.4.1 Ethyl Formate Pyrolysis

Simultaneous measurements of H2O and CO concentration time-histories are plotted

together in Figure 5.7 at temperatures between 1314 and 1629 K, pressures between 1.48

and 1.69 atm. Nearly equal formation rate (CO is slightly faster) is observed for H2O and

CO during EF pyrolysis over the entire temperature range, except for the case at the

highest temperature (1629 K). As EF decomposition is dominated by the concerted

unimolecular elimination to produce ethylene and formic acid, the measured 1:1 of

H2O/CO ratio is good evidence for the dehydration reaction of formic acid (HCOOH =

H2O + CO). Note that the H2O/CO ratio is less than 1 (actually 0.87) at the highest

temperature (1629 K) when these species time-histories reach their plateau levels at long

times. Such over-production of CO relative to H2O can be attributed to the added

competing EF decomposition channel (EF = CO + C2H5OH), which is more pronounced

at higher temperatures. Additionally, the product fractional yield shown in Figure 5.3(a)

reveals that CO2 is a minor product during EF pyrolysis with the CO2/H2O ratio between

0.05-0.07.

75

Figure 5.7 Measured H2O and CO concentration time-histories during the pyrolysis of ethyl formate.

Table 5.2 EF pyrolysis submechanism; cm3/mol/sec/cal units.

Reaction A n Ea RefEF=HCOOH+C2H4 1.00E+13 0.00 5.00E+04 [75]EF=CO+C2H5OH 1.15E+13 0.00 5.87E+04 [65]EF+H=EFp+H2 1.88E+05 2.80 6.28E+03 [75]EF+OH=EFp+H2O 1.05E+10 1.00 1.58E+03 [75]EF+CH3=EFp+CH4 1.29E+12 0.00 1.16E+04 [75]EFP=C2H4+OCHO 1.34E+13 -0.40 2.46E+04 [75]EF+H=EFs+H2 3.25E+05 2.40 4.47E+03 [75]EF+OH=EFs+H2O 1.16E+07 1.60 -3.50E+01 [75]EF+CH3=EFs+CH4 3.98E+11 0.00 9.50E+03 [75]EFs=CH3CHO+HCO 4.17E+15 -0.90 1.40E+04 [75]EF+H=EFf+H2 6.50E+05 2.40 4.47E+03 [75]EF+OH=EFf+H2O 2.33E+07 1.60 -3.50E+01 [75]EF+CH3=EFf+CH4 1.51E+00 3.50 5.48E+03 [75]C2H5+CO2=EFf 4.76E+07 1.50 3.74E+04 [75]C2H5O+CO=EFf 1.55E+06 2.00 5.73E+03 [75]EFp+H=EF 1.00E+14 0.00 0.00 [75]EFs+H=EF 1.00E+14 0.00 0.00 [75]EFf+H=EF 1.00E+14 0.00 0.00 [75]OCHO+C2H5=EF 1.00E+12 0.00 0.00 [75]HCO+C2H5O=EF 1.00E+12 0.00 0.00 [75]CH3+CH2OCHO=EF 1.00E+12 0.00 0.00 [75]HCOOH+M=CO+H2O+M 3.50E+16 0.00 5.40E+04 [93]a

HCOOH+M=CO2+H2+M 4.90E+15 0.00 5.70E+04 [93]a

HCO+OH = HCOOH 1.00E+14 0.00 0.00 [94]HCOOH+OH=CO2+H+H2O 2.62E+06 2.06 9.16E+02 [94]HCOOH+OH=CO+OH+H2O 1.85E+07 1.51 -9.62E+02 [94]HCOOH+H=CO2+H+H2 4.24E+06 2.10 4.86E+03 [94]HCOOH+H=CO+OH+H2 6.03E+13 -0.35 2.98E+03 [94]a The total rate is increased by a factor of 3.

Saito et al. [93] measured the branching ratio of formic acid decomposition in a

shock tube by monitoring the time-resolved IR radiation from CO (4.63 m) and CO2

76

(4.23 m). The branching ratio recommended by Saito et al. [93] was adopted in the

current model but the authors of the present work found it necessary to increase the two

decomposition rates by up to a factor of 3 for the best fit to the experimental data.

Although some independently corroborated experimental data exist for some conditions,

an overall large spread remains across all previous studies [91,93,97]. Rate constants for

the H-atom abstractions of formic acid are taken directly from the dimethyl ether (DME)

reaction kinetics by Fischer et al. [94], as formic acid is also an intermediate species in

DME oxidation. Figure 5.8 compares the measured H2O, CO and CO2 time-histories with

the model predictions using the current EF mechanism (Table 5.2). Excellent agreement

can be seen between the measured H2O and CO time-histories and simulations over the

entire temperature range. CO2 is also predicted to be a minor product as shown in Figure

5.8(c), in good agreement with the measurements.

(a)

(b)

77

(c)

Figure 5.8 Comparison of the measured (a) CO, (b) H2O and (c) CO2 concentration time-histories with the

model predictions during the pyrolysis of 2000 ppm EF in argon: solid line, measurement; dashed line,

simulation in this study.

(a) ROP (b) Sensitivity

Figure 5.9 ROP and sensitivity analyses of CO: 2000 ppm EF/Ar, 1500 K and 1.5 atm.

Rate of production (ROP) and sensitivity analyses of CO are plotted in Figure 5.9.

The ROP analysis (Figure 5.9(a)) indicates that CO is largely produced by the

dehydration reaction of formic acid (HCOOH = CO + H2O), which originates from EF

decomposition via a six-centered transition state. Hence, CO exhibits strong sensitivity to

EF and formic acid decomposition reactions as illustrated in Figure 5.9(b). However, EF

is completely consumed so quickly (within 15 s at 1500 K) that the CO sensitivity to EF

reactions is only pronounced at the very early times. Very similar performance of H2O

sensitivity can be seen in Figure 5.10, as H2O and CO are mainly produced from the

formic acid dehydration reaction. However, one of the EF unimolecular decomposition

78

reaction (EF = C2H5OH + CO, Rxn. C in Figure 5.9(b)) shows up in the early-time CO

sensitivity but is negligible in the H2O sensitivity plot. Therefore, this EF decomposition

channel accounts for the slight difference between CO and H2O concentration time-

histories measured behind reflected shock waves.

CO2 is observed to be a minor product during the pyrolysis of EF, which has also

been well captured by the current kinetic model. ROP analysis indicates that CO2 is

largely produced by the decarboxylation reaction of formic acid (HCOOH = CO2 + H2).

However, the formic acid decomposition favors the dehydration pathway with a H2O+CO

branching ratio of more than 0.9 under the shock tube conditions. The CO2 sensitivity

analysis shown in Figure 5.11 indicates that the CO2 mole fraction is dominantly

sensitive to those two competing pathways of formic acid unimolecular decomposition.

Figure 5.10 H2O sensitivity: 2000 ppm EF/Ar, 1500 K and 1.5 atm.

Figure 5.11 CO2 sensitivity: 2000 ppm EF/Ar, 1500 K and 1.5 atm.

79

5.4.2 Ethyl Acetate Pyrolysis

Current experimental results reveal that the CO production rate is much slower

compared to H2O and CO2 during EA pyrolysis, as shown in Figure 5.2(b). Additionally,

the CO time-histories during the pyrolysis of EF, EA and EP at a temperature near 1450

K are plotted in Figure 5.12 for comparison. The CO formation rate during EA pyrolysis

was measured to be only 0.5 ppm/s in the first 500 s, which is 9 times slower than EF

and more than 6 times (considering the 30 K difference) slower than EP, respectively.

Hence, it is of interest to investigate the kinetic interpretation of this experimental

observation.

Figure 5.12 Comparison of the measured CO concentration time-histories during the pyrolysis of EF, EA

and EP; pressure near 1.5 atm, fuel concentration 2000 ppm.

The kinetic mechanism specific to EA pyrolysis (Table 5.3) has been built by

considering the current experimental results and a bibliographic study on ethyl acetate

and acetic acid. The EA submechanism is taken from Westbrook et al. [75] without

modification except for the rate constant of EA bond-fission reaction (EA = CH3 +

C2H5OCO), which is increased by a factor of two. Rates for this reaction used in the core

EA kinetic mechanism [75] were estimated from the reverse radical-radical

recombination reaction while ignoring the barrier height, and thus the adjustments here

are justified. The submechanism of acetic acid is one of the core subsets for the EA

pyrolysis model, since acetic acid is the major intermediate produced from the initial EA

unimolecular elimination (EA = CH3COOH + C2H4). Leplat and Vandooren [95]

recently performed numerical and experimental study of the combustion of acetic acid in

three CH3COOH/O2/Ar low-pressure premixed flames. Considering the different

80

experimental conditions, their model is adopted here as the initial estimate for modeling

our shock tube data.

Table 5.3 EA pyrolysis submechanism; cm3/mol/sec/cal units.

Reaction A n Ea RefEA=CH3COOH+C2H4 2.00E+13 0.00 5.00E+04 [75]C2H5O+CH3CO=EA 3.00E+13 0.00 0.00 [75]C2H5+CH3CO2=EA 3.00E+13 0.00 0.00 [75]C2H5OCO+CH3=EA 6.00E+13 0.00 0.00 [75]a

EA+H=EAp+H2 1.88E+05 2.80 6.28E+03 [75]EA+OH=EAp+H2O 1.05E+10 1.00 1.58E+03 [75]EA+CH3=EAp+CH4 1.29E+12 0.00 1.16E+04 [75]EAp=C2H4+CH3CO2 1.34E+13 -0.40 2.46E+04 [75]EA+H=EAs+H2 3.25E+05 2.40 4.47E+03 [75]EA+OH=EAs+H2O 1.16E+07 1.60 -3.50E+01 [75]EA+CH3=EAs+CH4 3.98E+11 0.00 9.50E+03 [75]EAs=CH3CHO+CH3CO 4.17E+15 -0.90 1.40E+04 [75]EA+H=EAm+H2 6.50E+05 2.40 2.58E+03 [75]EA+OH=EAm+H2O 1.40E+10 0.50 6.30E+01 [75]EA+CH3=EAm+CH4 1.51E-10 6.40 8.93E+02 [75]CH2CO+C2H5O=EAm 1.00E+13 0.00 0.00 [75]EAp+H=EA 1.00E+13 0.00 0.00 [75]EAs+H=EA 1.00E+13 0.00 0.00 [75]EAm+H=EA 1.00E+13 0.00 0.00 [75]CH3COOH=CH4+CO2 7.08E+13 0.00 7.46E+04 [98]CH3COOH=CH2CO+H2O 4.47E+14 0.00 7.98E+04 [98]HOCO+CH3=CH3COOH 1.20E+12 0.00 0.00 b

CH3COOH+H=CH2COOH+H2 8.40E+07 2.00 7.70E+03 [95]CH3COOH+H=CH3CO2+H2 5.55E-23 10.6 -4.46E+03 [95]CH3COOH+OH=CH2COOH+H2O 1.29E+10 1.10 1.81E+03 [95]CH3COOH+OH=CH3CO2+H2O 2.40E+11 0.00 -4.00E+02 [95]CH3COOH+CH3=CH2COOH+CH4 6.60E+11 0.00 2.78E+03 [95]CH3COOH+CH3=CH3CO2+CH4 6.11E+00 3.57 7.72E+03 [95]CH2+CO(+M)=CH2CO(+M) k∞

k0

8.10E+112.69E+33

0.50-5.11

4.51E+037.09E+03 [60]

= 0.5907, T*** = 2.75E+02, T* = 1.226E+06, T** = 5.185E+03CH2CO+H=CH3+CO 1.10E+13 0.00 3.40E+03 [83]CH2CO+H=HCCO+H2 5.00E+13 0.00 8.00E+03 [83]c

CH2CO+OH=HCCO+H2O 1.00E+13 0.00 2.00E+03 [83]CH2CO+OH=CH2OH+CO 2.00E+12 0.00 -1.01E+03 [83]CH2+CH2CO=C2H4+CO 1.60E+14 0.00 0.00E+00 [83]CH2CO+CH3=C2H5+CO 5.00E+12 0.00 0.00E+00 [95]a A-factor increased by a factor of 2.b Estimated with the best fit to the experimental data.c A-factor divided by a factor of 4.

81

(a)

(b)

(c)

Figure 5.13 Comparison of the measured (a) CO, (b) H2O and (c) CO2 concentration time-histories with the

model predictions for 2000 ppm EA/Ar: solid line, measurement; dashed line, simulation.

82

The unimolecular decomposition of acetic acid has been investigated over the

temperature range of 1300-1950 K in a shock tube [98]. Based on the experimental

observation of the principal products such as ketene, H2O, CO2 and methane, Mackie and

Doolan [98] proposed that acetic acid decomposes mainly via the dehydration

(CH3COOH = CH2CO+H2O) and decarboxylation (CH3COOH = CH4+CO2) pathways.

The rate constants recommended by Mackie and Doolan [98] were used in this study

without modification. Another radical-related channel for acetic acid decomposition

(CH3COOH = CH3 + HOCO) is also considered in the modeling to achieve a best fit to

the experimental data. The ketene reactions and other submechanims are all based on the

core C1-C4 models in the Metcalfe et al. [83] mechanism. All the measured H2O, CO and

CO2 concentration time-histories are plotted in Figure 5.13, along with simulations using

the current EA mechanism. Simulations are in relatively good agreement with

measurements over all the experimental conditions, though differences remain that merit

further adjustments in the mechanism.

The ROP analysis shown in Figure 5.14(a) indicates that CO is mainly produced

from ketene, through ketene unimolecular decomposition (CH2CO = CH2 + CO, Rxn. D

in Figure 5.14(a)), and through the ketene bimolecular reactions with CH3, H and OH

radicals (CH2CO + CH3 = C2H5 + CO, Rxn. A in Figure 5.14(a); CH2CO + H = CH3 +

CO, Rxn. B in Figure 5.14(a)); and CH2CO + OH = CH2OH + CO, Rxn. C in Figure

5.14(a)). These reactions also appear in the CO sensitivity plot as illustrated in Figure

5.14 (b). Nearly all EA takes the unimolecular elimination (EA = C2H4 + CH3COOH,

Rxn. A in Figure 5.14 (b)) to produce the intermediate species acetic acid, which quickly

decomposes via two competing pathways to CO2+CH4 (decarboxylation) and

H2O+CH2CO (dehydration). It should be noted that the existence of methyl group in

acetic acid leads to the formation of ketene during the dehydration process instead of the

direct formation of CO during the dehydration of formic acid. The initial EA

decomposition involves little production of radicals, since the major EA initiation

reaction (EA = C2H4 + C2H5COOH) and subsequent acetic acid decomposition

(C2H5COOH = CO2 + CH4, C2H5COOH = H2O + CH2CO) are all concerted molecular

reactions. Hence, the CO formation is constrained by the ketene decomposition rate and

the size of CH3, H and OH radical pool in the system.

83

(a) ROP (b) SensitivityFigure 5.14 (a) ROP and (b) sensitivity analyses (using the current EA mechanism) of CO during the

pyrolysis of 2000 ppm EA/Ar at 1500 K and 1.5 atm.

Different from formic acid decomposition (major intermediate during EF pyrolysis)

which significantly favors the dehydration channel, acetic acid proceeds through

dehydration and decarboxylation decomposition channels with almost equal rate. It

explains the formation of much larger amount of CO2 during the pyrolysis of EA

compared to that of EF (differs by a factor of six or more). As expected, the CO2

sensitivity plotted in Figure 5.15 indicates that the CO2 mole fraction is dominantly

sensitive to the decarboxylation reaction (CH3COOH = CO2 + CH4) of acetic acid. Much

weaker sensitivity can be seen to EA unimolecular decomposition (EA = CH3COOH +

C2H4 and EA = C2H5 + CH3CO2) at the very early times and the H-atom abstraction

reaction of acetic acid (CH3COOH + OH = CH2COOH + H2O) at the long times.

Considering the remaining underprediction of CO2 at the lowest temperature (1492 K),

the rate constants assigned to these reactions may require further investigation. These

reactions also appear in the H2O sensitivity as illustrated in Figure 5.16. The H2O

concentration is sensitive to the branching ratio of acetic acid unimolecular elimination,

and the H-atom abstraction of acetic acid by hydroxyl radical.

84

Figure 5.15 CO2 sensitivity during the pyrolysis of 2000 ppm EA/Ar at 1500 K and 1.5 atm.

Figure 5.16 H2O sensitivity during the pyrolysis of 2000 ppm EA/Ar at 1500 K and 1.5 atm.

5.4.3 Ethyl Propanoate Pyrolysis

The detailed kinetic mechanism by Metcalfe et al. [83] is used for the analysis of

ethyl propanoate pyrolysis. In general, the model fails to predict all the H2O, CO and CO2

concentration time-histories during the pyrolysis of EP behind reflected shock waves.

Figure 5.17 presents the measured product fractional yields at 1 ms, along with the model

predictions using the Metcalfe et al. [83] mechanism. At the highest temperature near

1600 K, the model significantly underpredicts the CO2 yield by a factor of 5, but

overpredicts the H2O and CO yield by 50% and 30%, respectively. These significant

discrepancies indicate that the EP decomposition pathways need to be revised in the

Metcalfe et al. [83] mechanism.

85

Figure 5.17 Measured (symbol-solid line) and simulated (dashed line, Metcalfe et al. [83]) CO, H2O and

CO2 yields for 2000 ppm EP/Ar mixture at 1 ms. Temperature: 1301-1580 K; pressure: 1.4-1.7 atm.

98.6%

50%

+C2H4

COCO2

46.3%

+H, OH32%

+H,OH65%

1.7%

37.5% -H,12.5% 50%

3.7% +H,OH5%

+OH2%

+OH, 2.3%

C2H4 +

Figure 5.18 Main reaction pathways for EP pyrolysis using the Metcalfe et al. [83] mechanism: 2000 ppm

EP/Ar, 1350 K, 1.5 atm, at t = 200s.

Figure 5.18 presents the main pathways for CO and CO2 production during the

pyrolysis of EP using the Metcalfe et al. [83] mechanism at 1350 K and 1.5 atm; in this

scheme, the thickness of the arrow is proportional to the importance of the chemical

pathway. At t = 200 s, more than 98 percent of EP decomposes to ethylene and

propanoic aicd (EP = C2H5COOH + C2H4) through a six-center transition state.

86

Subsequent reactions of propanoic acid include three decomposition pathways. The first

pathway (32%) involves the H-atom abstraction to produce a CH2CH2COOH radical,

followed by the -scission to yield one ethylene and one HOCO radical. The unstable

HOCO radical continues to decompose to CO2 and CO with the CO2/CO ratio

approximately 0.08 in the Metcalfe et al. [83] mechanism, indicating nearly all HOCO is

converted to CO. The second pathway (65%) involves the H-atom abstraction of

propanoic acid to produce a CH3CHCOOH radical, followed by the -scission to yield

methyl ketene and hydroxyl radical, and the H-atom abstraction to produce a propenoic

acid molecule. In the Metcalfe et al. [83] model, methyl ketene is relatively stable under

current experimental conditions; only 4% of methyl ketene reacts with OH radical to

yield CO and CO2 at t = 200s. Finally, a small amount of propanoic acid (less than 2%)

takes the third pathway to produce the CH2COOH radical through bond cleavage,

followed by the -scission to produce ketene. Therefore, the CO2 yield is determined

mainly by the branching ratios of HOCO radical and methyl ketene decomposition.

As discussed before, the Metcalfe et al. [83] mechanism adopted a CO2 branching

ratio less than 0.1 for HOCO decomposition over the temperature range of 1000-2000 K.

HOCO radical is treated as an important intermediate in the CO + OH reaction system

[99–101]. The rate constant for reaction HOCO = CO + OH can be obtained from the

reverse reaction with the well-known equilibrium constant [99,101]. However, no study

can be found in terms of the branching ratios of those two HOCO radical decomposition

channels: HOCO = CO2 + H and HOCO = CO + OH. Interestingly, another radical

CH3OCO, showing similar potential energy surface to that of HOCO [102], has been

thoroughly studied recently as it is a crucial radical formed during methyl butanoate

decomposition. Huynh et al. [103] calculated the rate constants of CH3OCO

decomposition (CH3OCO = CO2 + CH3 and CH3OCO = CO + CH3O) using the RRKM

theory with corrections from tunneling, hindered rotation and variational treatments. The

branching ratio for the CO2 channel was calculated to be 0.7-0.9 over the temperature

range of 1000-2000 K [103]. Additionally, some preliminary results from the recent

calculation by John Barker [104] using the MultiWell code give the CO2+H branching

ratio to be 0.4-0.5 for HOCO decomposition over the temperature range of 1300-1600 K

87

and near 1 atm. Hence, in this study the branching ratio of the CO2+H channel for HOCO

decomposition is modified to be between 0.4-0.5 over the temperature range of 1300-

1600 K. Without any experimental evidence, an uncertainty factor of two or more are still

assigned to the rate constants of these reactions.

The rate constant for H-atom abstraction of propanoic acid (C2H5COOH + H =

CH2CH2COOH + H2) is increased by a factor of three to improve the agreement with

experimental data. The proposed changes to the rate constant of H-abstraction of

propanoic acid are justified because the propanoic acid submechanism in the Metcalfe et

al. model is constructed based on n-heptane and iso-octane rates [83]. According to the

recent flat flame study by Yang et al. [85], the rate constants for H-atom abstraction of

EP were modified for better agreement with the measured composition of reaction

intermediates in the low-pressure flames. The rate constants for these reactions were

modified in general by a factor of 2-5 by Yang et al. [85]. Therefore, the Metcalfe et al.

[83] mechanism is also updated with the new rate constants from Yang et al. [85]. All the

reactions with rate constants modified in the Metcalfe et al. [83] mechanism are

summarized in Table 5.4.

Table 5.4 Reaction rate constants modified in the Metcalfe et al. [83] mechanism; cm3/mol/sec/cal units.

Reaction A n Ea RefEP+H=EP3J+H2 1.33E+06 2.54 6.76E+03 [85]EP+H=EP2J+H2 5.04E+13 0 7.30E+03 [85]EP+H=EPEJ+H2 3.25E+05 2.4 2.58E+03 [85]EP+H=EPMJ+H2 1.88E+05 2.54 6.76E+03 [85]EP+OH=EP3J+H2O 1.06E+10 0.97 1.59E+03 [85]EP+OH=EP2J+H2O 2.30E+10 0.51 6.30E+01 [85]EP+OH=EPEJ+H2O 2.29E+10 0.51 6.30E+01 [85]EP+OH=EPMJ+H2O 1.05E+10 0.97 1.59E+03 [85]C2H5COOH+H=CH2CH2COOH+H2 1.50E+06 2.54 6.76E+03 [83]a

HOCO=CO+OH 4.56E+26 -5.12 2.76E+04 [100][102]HOCO=CO2+H 1.07E+36 -8.11 2.90E+04 [100][102]

a A-factor increased by a factor of 2.

88

(a)

(b)

(c)

Figure 5.19 Comparison of measured (a) H2O and (b) CO2 and (c) CO concentration time-histories with the

model predictions during the pyrolysis of 2000 ppm EP/Ar. Solid line, current measurement; dash-dot line,

simulation using the Metcalfe et al. [83] mechanism; dashed line, simulation using the modified Metcalfe et

al. mechanism.

89

Figure 5.19 presents the comparison of the measured H2O, CO and CO2 time-

histories and the model predictions using the original and modified Metcalfe et al. [83]

mechanism. The modifications proposed in this study significantly improve the

agreement between measurements and simulations. The modified EP model captures the

product yields of H2O, CO and CO2 fairly well over the entire temperature range, but the

early-time formation of CO and CO2 rate is still underpredicted.

ROP and sensitivity analyses for CO are illustrated in Figure 5.20, conducted at

1500 K and 1.5 atm for 2000 ppm EP in argon. The ROP analysis indicates that CO is

largely produced through the HOCO decomposition (HOCO = CO + OH) at the early

times and methyl ketene decomposition (CH3CHCO + H = C2H5 + CO, CH3CHCO +

CH3 = iC3H7 + CO and CH3CHCO + OH = sC2H4OH + CO) at the long times. Therefore,

the CO concentration must be sensitive to those reactions building radical pools.

Sensitivity analysis shown in Figure 5.20(b) supports the ROP interpretation. The early-

time CO concentration is strongly sensitive to the initial EP unimolecular elimination (EP

= C2H5COOH + C2H4) with negative effect and bond fission reaction (EP = C2H5CO2 +

C2H5) with positive effect. Subsequent decomposition reactions of C2H5CO2 and C2H5

radicals both produce the H atoms. CO also shows sensitivity to propanoic acid bond

fission reaction (C2H5COOH = CH3 + CH2COOH, followed by CH2COOH = CH2CO +

OH), as this is a radical branching reaction producing both CH3 and OH radicals. Large

amounts of methyl ketene and HOCO radicals are produced by the -scission of

CH3CHCOOH and CH2CH2COOH intermediates, both primarily originating from the

H-atom abstractions of propanoic acid (C2H5COOH + H = CH3CHCOOH + H2 and

C2H5COOH + H = CH2CH2COOH + H2). These reactions exhibit high sensitivity at the

longer times of CO time-histories as shown in Figure 5.20(b). All of these reactions

involving methyl ketene, propanoic acid and EP may contribute to the discrepancies

between the current measurements and simulations. However, it is not clear which of

these reactions must be modified, and further study is necessary to improve the EP

kinetic mechanism.

90

(a) ROP (b) SensitivityFigure 5.20 CO (a) ROP and (b) sensitivity analyses using the modified Metcalfe et al. [83] mechanism:

2000 ppm EP/Ar, 1500 K and 1.5 atm.

Figure 5.21 CO2 sensitivity analysis using the modified Metcalfe et al. [83] mechanism: 2000ppm EP/Ar,

1500 K and 1.5 atm.

The modified Metcalfe et al. [83] mechanism adequately captures the CO2 yield at

long times, but still underpredicts the early-time CO2 concentrations; see Figure 5.19(b).

Figure 5.21 presents the CO2 sensitivity at 1500 K and 1.5 atm for 2000 ppm EP/Ar

mixture. The branching ratio of the H-atom abstraction of propanoic acid to produce

CH3CHCOOH and CH2CH2COOH determines the final CO2 yield, as these two

competing pathways show strong sensitivity over the entire CO2 time-histories but with

opposite effect. Similar to the CO sensitivity plot shown in Figure 5.20(b), the EP

unimolecular decomposition (EP = C2H5COOH + C2H4 and EP = C2H5CO2 + C2H5) and

propanoic acid bond fission (C2H5COOH = CH3 + CH2COOH) play significant roles in

91

determining the early-time CO2 concentration. As evident from the discrepancies

between simulations and experimental data, further refinements to the Metcalfe et al. [83]

mechanism are needed. Because of the large number of reaction pathways, a final

recommendation for all the key individual reaction rates in the EP decomposition

mechanism cannot yet be made. The direct measurement of certain reaction rates,

however, is feasible and may provide a worthwhile research path. High-level ab initio

calculations are also recommended to reduce the uncertainties in the rate constants for EP

and propanoic bond fission, and methyl ketene decomposition reactions.

In summary, extensive high-quality, multi-species time-history data are presented

for EP pyrolysis, providing unique evaluation and refinement of the existing kinetic

mechanism [83]. Sensitivity and ROP analyses indicate that species time-histories are

dependent on a complicated network of chemical reactions, many of which have not been

well studied. Strong evidence is presented indicating that reactions related to methyl

ketene bimolecular decomposition, propanoic acid and EP unimolecular elimination

significantly affect the CO and CO2 production. Further studies for modifications to these

rate constants are required to improve performance of the EP kinetic mechanism.

92

93

Chapter 6. Summary and Future Directions

6.1 Summary of Results

6.1.1 Mid-IR CO Sensor near 4.7 m

A QC-laser-based absorption sensor for CO and temperature in high-temperature

shock-heated gases was reported using the fundamental band of CO near 4.7 m. The

selected transitions, v” = 0, R(12), R(13), P(20) and v” = 1, R(21), R(22), P(14) were

successfully accessed by two different QC lasers. The spectroscopic parameters including

line-strengths and broadening coefficients 2Ar-CO were determined at room-temperature

(296 K) and high temperatures (1100-2000 K) and compared with literature values.

A scanned-wavelength direct absorption CO sensor using a single QC laser was first

validated for accurate measurements of temperature and CO concentration in a shock

tube. The sensor measured temperature at a scan rate of 2.5 kHz by comparing the

measured peak absorbance ratio of the line pair R(21) and R(12) with spectral

simulations, showing very good agreement (within 0.8%) with the calculated

temperatures at 1300-2200 K. A fixed-wavelength CO temperature sensor based on

transitions R(21) and P(20), accessed by two different lasers centered at 2194.46 cm-1 and

2059.91 cm-1, was also developed to provide in situ detection with faster time response.

Sensor validation was first demonstrated in a shock tube by measuring temperatures

(1200-1900 K) and CO concentrations of CO/H2/Ar mixtures with 1 MHz bandwidth.

The sensor was then applied to the shock tube study of the pyrolysis and oxidation of

methyl formate by measuring CO concentration and temperature time-histories to

illustrate its capability in chemical kinetic studies. The increased absorption strength in

94

this wavelength region provides opportunities for more sensitive and accurate combustion

measurements with shorter optical path length and lower CO concentration than was

possible using overtone band absorption.

6.1.2 Two-Line Thermometry for Multiphase Flows

A TDL absorption sensor with a time-response of 40 kHz has been presented for

accurate temperature measurements in pure gases and in an environment with significant

attenuation of incident light by aerosol scattering. The current sensor probes the R(28)

and P(70) absorption transitions of the v1+v3 band of CO2 near 2.7 m, which were

selected for sensitive temperature measurements over a wide range of temperatures (600-

1500 K). The fixed-center-wavelength WMS with 2f detection was used for calibration-

free measurements by normalization using the 1f signal. Experiments conducted in an

aerosol flow cell demonstrate that the sensor has the potential to measure gas temperature

accurately, even when the droplet scattering attenuates more than 99% of the incident

intensity.

The temperature sensor was first validated in non-reactive CO2/Ar gas mixtures in a

shock tube. Excellent agreement (within 1.5%) was found between the measured

temperatures (incident shock: 650-850 K, reflected shock: 1100-1500 K) and the

calculated values using shock jump equations. The sensor was then applied in successful

measurements of CO2 temperature in evaporating n-dodecane aerosols in an aerosol

shock tube. The temperature measured prior to the complete evaporation of the droplets

reflected the true temperature of the shock-heated test gas/aerosol mixture. Temperatures

were measured over the 520-1200 K range and varied less than 1.8% from the expected

values calculated using a laboratory-developed code. The WMS-2f/1f CO2 sensor

described in this study shows good potential for applications in a wide variety of rapidly-

varying, multi-phase environments.

6.1.3 Methanol and Methyl Formate Decomposition Study

Methanol and methyl formate thermal decomposition was studied experimentally by

measuring CH3OH and CO concentration time-histories behind reflected shock waves. A

95

quantitative absorption diagnostic was developed for measuring methanol concentration

at high temperatures using CO2 laser absorption spectroscopy at 9.676 m. The CO

measurement was made using a QC laser to access the R(13) transition line of the CO

fundamental band near 4.56 m. These two laser absorption diagnostics provided time-

resolved species concentration data with high signal-to-noise ratio even with a highly

diluted mixture (1000 ppm).

In the study of methanol decomposition, a comparison of the current measurements

with the model predictions from the Li et al. [17] mechanism, combined with the

sensitivity analysis, identified the need to include another methanol decomposition

channel CH2(S)+H2O (and associated CH2(S) reactions), in direct competition with the

CH3+OH channel. With the rate constants of methanol unimolecular elimination and H-

atom abstraction reactions modified with recent theoretical values from Jasper et al. [53]

and Meana-Pañeda et al. [63], good agreement was found between the modified Li et al.

[17] mechanism and the measured methanol and CO time-histories. However, in light of

the remaining small discrepancy in CO time-histories, we recommend further scrutiny of

the reactions involving hydroxymethyl and formaldehyde, especially the decomposition

reaction CH2OH → CH2O + H and the H-elimination reaction CH2OH +H → CH3 + OH.

In the study of methyl formate decomposition, the reaction rate constants for the

dominant MF decomposition channel (MF → CH3OH + CO) were studied behind

reflected shock waves by fitting the measured CO time-histories to the simulations using

the Dooley et al. mechanism [33]. Our measurement is in fairly good agreement with the

estimations from Dooley et al. [33] and the theoretical calculations by Peukert et al. [74].

These are the first high-temperature rate measurements in this kinetically-simple system.

Considering the remaining discrepancies between measurements and model predictions

especially at long times of the CO and methanol time-histories, we believe the model

uncertainties propagate to the secondary reactions such as methanol submechanism

and/or methyl formate H-abstraction reactions. Several model modifications were

recommended to further improve the agreement of the model predictions with the current

experimental results.

96

6.1.4 Ethyl Ester Decomposition Study

Quantitative measurements of CO, CO2 and H2O concentration time-histories were

carried out for ethyl formate, ethyl acetate, and ethyl propanoate pyrolysis using mid-IR

laser absorption techniques in a shock tube. More than 90% of oxygen balance was

achieved for all these esters when CO, CO2 and H2O time-histories reached the plateau

within the test time of 2 ms. These C3-C5 ethyl esters mainly take the unimolecular

elimination reaction to produce ethylene and the corresponding acid, i.e., formic acid for

EF, acetic acid for EA, propanoic acid for EP. Hence, the final product yields are mainly

determined by the submechanisms of these carboxylic acids. Detailed kinetic modeling

was performed to understand the experimental results. The EF and EA pyrolysis models

proposed in the current study can simulate the corresponding species time-histories fairly

well, while the detailed EP mechanism by Metcalfe et al. [83] needs to be further

modified to match our experimental data. Reaction pathway and sensitivity analyses

indicate that the branching ratios of HOCO radical decomposition and H-abstraction of

propanoic acid need to be modified, as the CO2 yield was significantly underpredicted in

the Metcalfe et al. [83] mechanism. Further experimental and theoretical studies of

certain reaction rates such as HOCO, ketene and methyl ketene decomposition are

recommended to reduce the uncertainties in the current kinetic models.

6.2 Recommendations for Future Work

6.2.1 Shock Tube Measurements of Reaction Rate Constants

The shock tube/laser diagnostic tool has been demonstrated to be powerful in

determining the rate constants of elementary reactions. Based on the conclusions in this

study, there are quite a few reactions playing important roles in the current kinetic

mechanisms but with rate constants not very well known. Taking HOCO chemistry as an

example, there are two competing pathways (HOCO = CO + OH, HOCO = CO2 + H) for

HOCO radical thermal decomposition. The branching ratio of HOCO decomposition is

the key to a lot of the modeling efforts on alkyl ester pyrolysis and possibly affecting

97

butanol combustion chemistry also. Large uncertainties (a factor of 2 or more) exist in the

theoretical calculations of these rate constants, and unfortunately no experimental data

can be found till now.

Current shock tube and mid-IR laser absorption provide the promising opportunity

to study these important reactions. First of all, benzoic acid can be used as a HOCO

radical precursor. At relatively high temperatures, benzoic acid dominantly takes the

following unimolecular decomposition to produce HOCO radical:

C6H5COOH ↔ C6H5 + HOCO, (1)

followed by the subsequent HOCO radical decomposition:

HOCO ↔ CO + OH, (2a)

and

HOCO ↔ CO2 + H. (2b)

Therefore, rate constants for reactions (2a) and (2b) can be determined by monitoring CO

and CO2 concentration time-histories behind reflected shock waves. Due to the high

sensitivity of the current mid-IR laser absorption diagnostics, very dilute fuel

concentration can be used to eliminate the interference from secondary reactions.

6.2.2 Multi-Species Measurements in Large Oxygenates and Blends

Most of the previous shock tube studies have been performed on small biodiesel

surrogates due to their simpler chemical structure and experimental availability for

conventional gas-phase shock tubes. However, the real biodiesel blends are usually

composed of large FAMEs, such as methyl oleate (C19H36O2), methyl linoleate

(C19H34O2), methyl linolenate (C19H32O2), methyl palmitate (C17H34O2), and methyl

stearate (C19H38O2). Experimental data for these large oxygenate compounds would be

extraordinarily important for the kinetic modeler in terms of model validation and

refinement. Hence, it is natural to come up with the idea of extending the multi-species

diagnostic strategy to these large methyl esters in future.

Studying larger surrogates in conventional gas-phase shock tubes has proven

extremely difficult due to these fuels' low vapor pressures and associated decomposition

issues in preparing gaseous reactant mixtures. The aerosol shock tube technique at

98

Stanford University significantly extends the range of fuels observable behind reflected

shocks [105]. Ignition delay time measurements for methyl oleate and methyl linoleate

have been recently conducted in our laboratories [106]. Several of the important IR-laser-

based diagnostics recently developed in our lab and the potential new mid-IR diagnostics

in need are summarized in Table 6.1 and Table 6.2, respectively. Combined with the UV

diagnostics for radicals (OH and CH3), those measured species time-history and rate data

provide additional opportunity to test and validate large reaction mechanisms and refine

their component sub-mechanisms.

Furthermore, while formulating surrogate kinetic models by mixing different single

component submechanisms, the " cross-term " reactions may become important and the

presence of each component can affect the oxidation chemistry of other [107,108]. Co-

oxidation between reactants occurs quasi-exclusively via the radical pool. Therefore,

shock tube multi-species measurements in multi-component blends or the real biodiesel

would serve as one of the most stringent kinetic validation targets.

Table 6.1 Stanford IR laser diagnostics for combustion gases

Species Wavelength[m]

Line-Strength (296K)[cm-2atm-1]

Laser System

H2O 2.5 5.80 DFBCO2 2.7 1.00 DFBC2H2 3.0 5.90 DFBCH4 3.4 3.30 DFGCO2 4.3 87.0 QCLCO 4.6 11.0 QCLNO 5.2 0.60 QCLCH3OH 9.6 0.50 CO2 LaserC2H4 10.5 2.10 CO2 Laser

Table 6.2 New species and potential diagnostics in future

Species Wavelength[m]

Laser System

propene 10.95 QCLn-butene 10.9 QCLi-butene 11.3 QCLallene 11.5 QCL

99

6.2.3 Kinetics of Oxygenated Fuel Thrust

Nowadays chemical kinetic models are often developed in a hierarchical manner

starting with small molecules with only a few atoms, and this approach has proved to be

effective and powerful for the modeling of large hydrocarbon fuels [57,109–111]. The

global model performance heavily relies on the accuracy of these mechanism thrusts. The

kinetic studies carried out in this thesis on the thermal decomposition of oxygenates have

elucidated many areas relating to the thrusts (C1-C4) of the oxygenated fuels. Some of

these foundational submechanisms have received little kinetic attention due to their weak

sensitivity in hydrocarbon fuels, but may serve as essential model components in

oxygenated fuels.

In the current work, the thermal decomposition of ethyl esters is a representative

example showing the need for further refinements of the model thrusts. During the

pyrolysis of ethyl acetate and ethyl propanoate, the composition of the final products

depends highly on the reaction pathways of intermediates such as HOCO radical, ketene

and methyl ketene. However, the large uncertainties existing in the current

submechanisms for these oxygenates result in poor model predictions. In the case of

methanol pyrolysis, methanol is consumed mainly through H-atom abstraction to produce

CH2OH radical, which subsequently decomposes to formaldehyde via -scission or

reacts with H-atom to produce CH3 and OH. The current submechanisms involving

CH2OH and formaldehyde need further scrutiny due to the discrepancy between the

experimental results and the model predictions. When extended to the oxidation case, one

of the major consumption pathway for the oxygenates with the ester moiety is the

abstraction of the H atom from the CH2 group adjacent to the C=O function [112]. The

H-atom abstraction from these oxygenates produces radical species that react primarily

via -scission, leading to stable intermediates (ketenes, R'R"C=C=O) and smaller radicals

(methoxy/ethoxy). Very few experimental and theoretical studies on these species can be

found in the literature. Therefore, kinetics of these oxygenate thrusts would be of

immediate future interest and several important reactions can be measured using the

shock tube/laser diagnostic strategy.

100

101

Appendix A: Ethylene and Methanol

Diagnostics using CO2 Gas Laser

A.1 Ethylene Diagnostic at 10.532 m

Ethylene is a stable intermediate species and dominant alkene formed by fuel

fragmentation processes during the oxidation and pyrolysis of large alkanes [32,113].

Formation of ethylene and its subsequent reactions to acetylenic derivatives are also

involved in the formation and growth of polycyclic aromatic hydrocarbons (PAHs), the

most likely precursors to soot [114]. It follows that analysis of a wide variety of

hydrocarbon kinetic mechanisms would benefit from a sensitive and quantitative

diagnostic for ethylene concentration time-histories in shock tube experiments.

A.1.1 Experimental

Ethylene was monitored in shock tube measurements using CO2 laser absorption at

10.532 m. Absorption is due to the strong Q-branch of the v7 ethylene band, which has a

strong overlap with the P14 line of the CO2 laser transitions associated with the (0 0 1) to

(1 0 0) vibrational levels. We utilized a grating-tuned CO2 gas laser (Model Lasy-4G,

Access Laser Co.) with 230 mW output. The CO2 transition is primarily Doppler

broadened with a full width at half maximum (FWHM) of less than 100 kHz (i.e. 3×10-6

cm-1). The CO2 emission line was well-identified by passing a portion of the laser output

(through a beam splitter) into a mid-IR wavemeter (Bristol 721) and observed to be stable

over hours. The HITRAN database indicates that the absorption features of ethylene near

10.5 m have broadening coefficients on the order of ~0.1 cm-1atm-1 [13]. Thus, the CO2

laser emission at the P14 line can be considered monochromatic. New TE-cooled IR

102

photovoltaic detectors (Vigo Systems, PVM-2TE-10.6) with large linear dynamic range

were implemented. With this new detection system, we have found that the detection

noise can be reduced to <0.3% (typically over 2 ms, shock tube test time) even without

CMR. A schematic of the experimental setup is illustrated in Figure A.1.

Figure A.1 Schematic of CO2 laser diagnostic in shock tube measurements; ND: neutral density filter, NBP:

narrow bandpass filter.

A.1.2 High-Temperature Ethylene Absorption Cross-Section

For quantitative measurements of species concentration, the absorption cross-section

must be characterized by measuring the absorbance under known experimental conditions.

Gas mixtures of 1%, 0.5% and 0.25% C2H4/Ar were filled into the driven section of the

shock tube to the desired initial pressures (0.04-0.18 atm and 0.38-1.07 atm for the low-

pressure and the high-pressure shock tube facilities, respectively).

By filling helium into the driver section of the shock tube until the rupture of the

polycarbonate (in the low-pressure shock tube) or aluminum (in the high-pressure shock

tube) diaphragm, a shock wave was generated to compress and heat the test gas. The

C2H4/Ar mixtures were first shock-heated to 643-1075 K and 0.3-5.5 atm by the incident

shock waves and were then further heated and compressed to 1054-1959 K and 1.3-18.6

atm by the reflected shocks. Figure A.2 presents a representative example of the pressure

and the laser absorbance time-histories measured in the shock tube. The absorption cross-

section was determined by measuring the time-zero absorbance of the mixture

immediately behind shock waves. Note that at higher temperatures where C2H4 molecules

Signaldetector

ND

Shock tube

Iris

Referencedetector

NBP

Wavemeter

Wedgedbeam splitter

CO2 Laser

I0 I

Signaldetector

ND

Shock tube

Iris

Referencedetector

NBP

Wavemeter

Wedgedbeam splitter

Referencedetector

NBP

WavemeterWavemeter

Wedgedbeam splitter

CO2 LaserCO2 Laser

I0 I

103

decompose rapidly, the initial absorbance is extrapolated to time-zero to infer the

absorption cross-section.

The measured high-temperature ethylene cross sections are plotted in Figure A.3

with estimated uncertainties. Evident pressure dependence was observed at lower

pressures (0.3-1 atm); see incident shock data in Figure A.3. At pressures larger than 1

atm, however, the ethylene cross section was only weakly dependent on pressure across

the full temperature range of 643-1959 K. The absorption cross section can be modeled

as a product of two independent functions for T and P:

,T P T P Eqn. A-1

Since ethylene (T, P) at pressures larger than 1 atm is substantially sensitive to

temperature, the (T) factor can be first separated out from the measured ethylene cross

sections within a finite pressure range. Figure A.4 plots the measured ethylene cross

section (1.8-5.5 atm) as a function of temperature between 643 to 1895 K. Best fits of

experimental determinations of (T) in this pressure range yield the following double

exponential expression:

20 1 1 2 2,m /mol exp exp , 1.8 5.5 atmT a a T b a T b P Eqn. A-2

where a0 = 4.8, a1 = 383.7, and a2 = 103.5, in m2/mol; b1 = 183.0 and b2 = 378.8, in K.

This expression agrees with the measured values with an RMS deviation less than 1%.

By comparing the experimental data with the fitted (T) function, the pressure-

dependent factor (P) is then determined over the temperature range from 643 to 1959 K:

2

0.1

0.68 0.47 0.16 , 0.3 1.2 atm

0.82 0.2 , 1.2 18.6 atm

P P PP

P P

Eqn. A-3

According to Eqn. A-1,(T, P) is a direct product of (T) and (P) and is found to

have an RMS deviation of 1.4% for the data across the full range of pressure (0.3-18.6

atm) and temperature (643-1959 K). Note that the pressure dependence of the cross

section is seen to be much weaker than the temperature dependence. At a fixed pressure

of 1.5 atm, (T, P) decreased by 10% with 8% increase in temperature (from 1423 to

1547 K); however, at 1423 K, for a pressure change from 1.5-17 atm, (T, P) changed

by only 3.5%.

104

Figure A.2 Pressure and laser absorbance time-histories for a nonreactive mixture: 1% C2H4/Ar. Schlieren

spikes caused by the density gradient across the shock waves.

Figure A.3 Ethylene cross-sections (10.532m): 643-1959 K and 0.3-18.6 atm. Upper panel: measured

absorption cross section,meas; lower panel: comparisons of meas with fit calculated using (Eqn. A-1).

Figure A.4 Ethylene cross section (1.8-5.5 atm) as a function of temperature; best fit using (Eqn. A-2).

105

A.2 Methanol Diagnostic at 9.676 m

The methanol diagnostic briefly described in Chapter 4 uses the same CO2 gas laser

as that for ethylene diagnostic. The infrared spectra of gas-phase methanol reported by

Pacific Northwest National Laboratories (PNNL) are demonstrated in Figure A.5

between 2 and 12 m [115]. In this wavelength range, besides the relatively weak

absorption bands corresponding to the CH3‒ stretching (3.4 m), ‒OH stretching (2.7 m)

and CH3‒ bending (7.4 m), methanol has a very strong v8 fundamental absorption band

at 9.676 m associated with the C‒O stretching. Fortuitously, this C‒O stretching

frequency has a strong overlap with the P34 CO2 line in the (0 0 1) to (0 2 0) vibrational

level [58], providing an excellent strategy for measuring methanol time-histories in

combustion processes.

Figure A.5 IR spectra of gas-phase methanol at 298 K (from PNNL [115])

A.2.1 Methanol Absorption Cross-Section

The methanol absorption cross-section at room temperature (297 K) was measured

in a static cell with a path length of 25.4 cm. Methanol/Ar and methanol/N2 mixtures with

different mole fractions (xi =1.5%, 1% and 0.5%) were filled into the test cell to the

desired initial pressures (10-800 Torr). Figure A.6(a) shows a typical measured methanol

absorbance (normalized by xi and path length L) as a function of pressure. A constant

absorption cross-section of 76.6±0.3 m2/mol is inferred over the pressure range of 25-75

Torr. However, pressure dependence of the absorption cross-section is observed over a

106

wider range of pressure (10-200 Torr) as illustrated in Figure A.6(b). Comparisons of the

measured methanol absorption cross-section at 1 atm and 297 K with the previous studies

are also shown in Figure A.6(b) and summarized in Table A.1.

Shock tube was then used for methanol cross-section measurements at high

temperatures using the same method as discussed in Section A.1. The test mixtures were

first shock-heated to 665-1014 K and 0.4-0.8 atm by the incident shock waves and were

then further heated and compressed to 1126-1940 K and 1.4-2.7 atm by the reflected

shocks. The measured cross-section data were summarized and plotted in Figure A.7 with

estimated uncertainties. Under the current shock tube conditions, the methanol absorption

cross-section can be well expressed by an exponential function:

(T), m2/mol = 123.4×exp(-T, K/286.3)+4.66. Eqn. A-4

Considering the minimum detectivity of 0.1% absorbance using direct absorption

spectroscopy in shock tube experiments, this diagnostic strategy is capable of detecting

50-100 ppm level of methanol for a path length of 15 cm at 1200-1600 K and 1-2 atm.

(a) (b)

Figure A.6 Determination of methanol absorption cross-section (297 K) at 9.676 m: (a) measured

methanol absorbance as a function of pressure (10-70 Torr); (b) measured methanol cross-section as a

function of pressure (data for comparison at 1 atm are from Molina et al. [116], Sharpe et al. [115] and

Loper et al. [58]).

Table A.1 Methanol absorption cross-section (m2/mol) at 1 atm and 297 K.

Current study Molina et al.[116]

Sharpe et al.[115]

Loper et al.[58]

55.81.6 55.6 56.1 53.0

107

Figure A.7 Methanol cross-sections (m) measured over 665-1940 K and 0.4-2.7 atm. Curve fit is

given by (Eqn. A-4).

A.2.2 Two-Line Differential Absorption Measurement

The vibrational mode of the C-O group of methyl formate also absorbs at 9-10 m,

and this must be subtracted to obtain an accurate methanol measurement during the

pyrolysis of methyl formate. Hence, a second wavelength at 9.229 m was employed to

provide additional information of the interference absorption. As shown in Figure A.8,

the absorbance data at 9.676 m for 1% MF/Ar mixture at 1327 K and 1.5 atm consists of

absorption contributions from both methyl formate and methanol, with different

absorption cross-sections. In order to extract the mole fraction of methanol (also methyl

formate), a second wavelength at 9.229 m was employed to obtain additional

information. The absorption cross-sections for both species at these two different

wavelengths were measured and summarized in Figure A.9.

According to Beer’s law, the absorbance traces shown in Figure A.8 can be

expressed as follows:

1 1, MeOH MeOH 1, MF MF

2 2, MeOH MeOH 2, MF MF

x nL x nL

x nL x nL

Eqn. A-5

where 1 and 2 are the laser absorbances at 9.676 and 9.229 m, respectively. With all

the absorption cross-sections known, Eqn. (A-5) can thus be solved using simple algebra.

108

Figure A.8 Laser absorbance data for 1% MF/Ar mixture at 1327 K and 1.5 atm.

600 800 1000 1200 1400 16000

3

6

9

12

15

18

Methanol Methyl Formate

Abso

rptio

n C

ross

-sec

tion

[m2 /m

ol]

Temperature [K]

Wavelength = 9.676 m

(a)

600 800 1000 1200 1400 16000

3

6

9

12

15

18Ab

sorp

tion

Cro

ss-s

ectio

n [m

2 /mol

]

Temperature [K]

Methanol Methyl Formate

Wavelength = 9.229 m

(b)

Figure A.9 Measured absorption cross-sections of (a) methanol and (b) methyl formate at wavelengths of

9.676 and 9.229 m; P = 0.6-2.7 atm.

109

Reference

[1] International Energy Outlook 2013, Us Energy Information Administration, DOE (2003).

[2] M.G. Allen, Meas. Sci. Technol. 9 (1998) 545.

[3] H. Teichert, T. Fernholz, V. Ebert, Appl. Opt. 42 (2003) 2043.

[4] K. Kohse-Höinghaus, R.S. Barlow, M. Aldén, J. Wolfrum, Proc. Combust. Inst. 30 (2005)

89.

[5] R.K. Hanson, Proc. Combust. Inst. 33 (2011) 1.

[6] D.T. Cassidy, L.J. Bonnell, Appl. Opt. 27 (1988) 2688.

[7] R.M. Mihalcea, D.S. Baer, R.K. Hanson, Appl. Opt. 36 (1997) 8745.

[8] B.L. Upschulte, D.M. Sonnenfroh, M.G. Allen, Appl. Opt. 38 (1999) 1506.

[9] V. Weldon, J. O’Gorman, P. Phelan, J. Hegarty, T. Tanbun-Ek, Sensors Actuators B Chem.

29 (1995) 101.

[10] D.M. Sonnenfroh, M.G. Allen, Appl. Opt. 36 (1997) 3298.

[11] M.P. Arroyo, S. Langlois, R.K. Hanson, Appl. Opt. 33 (1994) 3296.

[12] J.T.C. Liu, G.B. Rieker, J.B. Jeffries, M.R. Gruber, C.D. Carter, T. Mathur, R.K. Hanson,

Appl. Opt. 44 (2005) 6701.

[13] HITRAN, http://www.cfa.harvard.edu/hitran/.

[14] F. Capasso, R. Paiella, R. Martini, R. Colombelli, C. Gmachl, T.L. Myers, M.S. Taubman,

R.M. Williams, C.G. Bethea, K. Unterrainer, H.Y. Hwang, D.L. Sivco, A.Y. Cho, A.M.

Sergent, H.C. Liu, E.A. Whittaker, Ieee J. Quantum Electron. 38 (2002) 511.

[15] A. Kosterev, G. Wysocki, Y. Bakhirkin, S. So, R. Lewicki, M. Fraser, F. Tittel, R.F. Curl,

Appl. Phys. B 90 (2008) 165.

[16] R.F. Curl, F. Capasso, C. Gmachl, A.A. Kosterev, B. McManus, R. Lewicki, M. Pusharsky,

G. Wysocki, F.K. Tittel, Chem. Phys. Lett. 487 (2010) 1.

[17] J. Li, Z. Zhao, A. Kazakov, M. Chaos, F.L. Dryer, J.J. Scire, Int. J. Chem. Kinet. 39 (2007)

109.

110

[18] R.M. Mihalcea, D.S. Baer, R.K. Hanson, Meas. Sci. Technol. 9 (1998) 327.

[19] M.E. Webber, J. Wang, S.T. Sanders, D.S. Baer, R.K. Hanson, Proc. Combust. Inst. 28

(2000) 407.

[20] J. Wang, M. Maiorov, D.S. Baer, D.Z. Garbuzov, J.C. Connolly, R.K. Hanson, Appl. Opt.

39 (2000) 5579.

[21] V. Ebert, H. Teichert, P. Strauch, T. Kolb, H. Seifert, J. Wolfrum, Proc. Combust. Inst. 30

(2005) 1611.

[22] X. Chao, J.B. Jeffries, R.K. Hanson, Meas. Sci. Technol. 20 (2009) 115201.

[23] R.K. Hanson, P.K. Falcone, Appl. Opt. 17 (1978) 2477.

[24] M. Schoenung, R.K. Hanson, Combust. Sci. Technol. 24 (1980) 227.

[25] J. Houston Miller, S. Elreedy, B. Ahvazi, F. Woldu, P. Hassanzadeh, Appl. Opt. 32 (1993)

6082.

[26] R. Barron-Jimenez, J.A. Caton, T.N. Anderson, R.P. Lucht, T. Walther, S. Roy, M.S.

Brown, J.R. Gord, Appl. Phys. B 85 (2006) 185.

[27] A.A. Kosterev, F.K. Tittel, R. Köhler, C. Gmachl, F. Capasso, D.L. Sivco, A.Y. Cho, S.

Wehe, M.G. Allen, Appl. Opt. 41 (2002) 1169.

[28] J. Vanderover, M.A. Oehlschlaeger, Appl. Phys. B 99 (2010) 353.

[29] X. Zhou, J.B. Jeffries, R.K. Hanson, Appl. Phys. B 81 (2005) 711.

[30] R.C. Millikan, D.R. White, J. Chem. Phys. 39 (1963) 3209.

[31] J.-P. Bouanich, C. Haeusler, J. Quant. Spectrosc. Radiat. Transf. 12 (1972) 695.

[32] D.F. Davidson, Z. Hong, G.L. Pilla, A. Farooq, R.D. Cook, R.K. Hanson, Proc. Combust.

Inst. 33 (2011) 151.

[33] S. Dooley, M.P. Burke, M. Chaos, Y. Stein, F.L. Dryer, V.P. Zhukov, O. Finch, J.M.

Simmie, H.J. Curran, Int. J. Chem. Kinet. 42 (2010) 527.

[34] R.J. Kee, F.M. Ruply, J.A. Miller, Chemkin- React. Des. Inc San Diego Ca (2012).

[35] R.K. Hanson, G.A. Pang, S. Chakraborty, W. Ren, S. Wang, D.F. Davidson, Combust.

Flame 160 (2013) 1550.

[36] J.T. Herbon, R.K. Hanson, D.M. Golden, C.T. Bowman, Proc. Combust. Inst. 29 (2002)

1201.

[37] Z. Hong, A. Farooq, E.A. Barbour, D.F. Davidson, R.K. Hanson, J. Phys. Chem. A 113

(2009) 12919.

[38] D.F. Davidson, D.R. Haylett, R.K. Hanson, Combust. Flame 155 (2008) 108.

111

[39] G.B. Rieker, J.B. Jeffries, R.K. Hanson, Appl. Phys. B 94 (2009) 51.

[40] A. Farooq, J.B. Jeffries, R.K. Hanson, Appl. Phys. B 90 (2008) 619.

[41] A. Farooq, J.B. Jeffries, R.K. Hanson, Appl. Phys. B 96 (2009) 161.

[42] F. Beyrau, A. Brauer, T. Seeger, A. Leipertz, Opt. Lett. 29 (2004) 247.

[43] A.R. Awtry, B.T. Fisher, R.A. Moffatt, V. Ebert, J.W. Fleming, Proc. Combust. Inst. 31

(2007) 799.

[44] J.M. Porter, J.B. Jeffries, R.K. Hanson, Appl. Phys. B 102 (2011) 345.

[45] H. Li, G.B. Rieker, X. Liu, J.B. Jeffries, R.K. Hanson, Appl. Opt. 45 (2006) 1052.

[46] C.F. Bohren, D.R. Huffman, Absorption and Scattering of Light by Small Particles, Wiley-

VCH, 1998.

[47] A.G. Thomas, Scatmech Polarized Light Scattering C++ Class Library. URL:

physics.nist.gov/Divisions/Div844/facilities/scatmech/html/index.htm.

[48] G.B. Rieker, J.B. Jeffries, R.K. Hanson, Appl. Opt. 48 (2009) 5546.

[49] M.A. Oehlschlaeger, D.F. Davidson, R.K. Hanson, J. Phys. Chem. A 108 (2004) 4247.

[50] A.E. Klingbeil, J.B. Jeffries, D.F. Davidson, R.K. Hanson, Appl. Phys. B 93 (2008) 627.

[51] R.G. Gordon, W. Klemperer, J.I. Steinfeld, Annu. Rev. Phys. Chem. 19 (1968) 215.

[52] E. L.Petersen, R. K.Hanson, (2012).

[53] A.W. Jasper, S.J. Klippenstein, L.B. Harding, B. Ruscic, J. Phys. Chem. A 111 (2007)

3932.

[54] A. Frassoldati, A. Cuoci, T. Faravelli, U. Niemann, E. Ranzi, R. Seiser, K. Seshadri,

Combust. Flame 157 (2010) 2.

[55] A. Demirbas, Energy Policy 35 (2007) 4661.

[56] E. Zervas, S. Poulopoulos, C. Philippopoulos, Fuel 85 (2006) 333.

[57] H.J. Curran, P. Gaffuri, W.J. Pitz, C.K. Westbrook, Combust. Flame 114 (1998) 149.

[58] G.L. Loper, A.R. Calloway, M.A. Stamps, J.A. Gelbwachs, Appl. Opt. 19 (1980) 2726.

[59] D.A. Sheen, H. Wang, Combust. Flame 158 (2011) 645.

[60] G.P. Smith, D.M. Golden, M. Frenklach, N.W. Moriarty, B. Eiteneer, M. Goldenberg, C.T.

Bowman, R.K. Hanson, S. Song, W.C. Gardiner, V. Lissianski, Z. Qin, GRI 3.0, URL:

http://www.me.berkeley.edu/gri_mech/.

[61] K. Spindler, H. Gg. Wagner, Berichte Bunsenges. Für Phys. Chem. 86 (1982) 2.

[62] T. Koike, M. Kudo, I. Maeda, H. Yamada, Int. J. Chem. Kinet. 32 (2000) 1.

[63] R. Meana-Pañeda, D.G. Truhlar, A. Fernández-Ramos, J. Chem. Phys. 134 (2011) 094302.

112

[64] W. Tsang, R.F. Hampson, J. Phys. Chem. Ref. Data 15 (1986) 1087.

[65] W. Ren, K.-Y. Lam, S.H. Pyun, A. Farooq, D.F. Davidson, R.K. Hanson, Proc. Combust.

Inst. 34 (2013) 453.

[66] J. Warnatz, in:, W.C.G. Jr (Ed.), Combust. Chem., Springer US, 1984, pp. 197–360.

[67] J.T. Jodkowski, M.-T. Rayez, J.-C. Rayez, T. Bérces, S. Dóbé, J. Phys. Chem. A 103 (1999)

3750.

[68] P.H. Cribb, J.E. Dove, S. Yamazaki, Combust. Flame 88 (1992) 169.

[69] J. Vandooren, P.J. Van Tiggelen, Symp. Int. Combust. 18 (1981) 473.

[70] D.L. Baulch, C.T. Bowman, C.J. Cobos, R.A. Cox, T. Just, J.A. Kerr, M.J. Pilling, D.

Stocker, J. Troe, W. Tsang, R.W. Walker, J. Warnatz, J. Phys. Chem. Ref. Data 34 (2005)

757.

[71] S. Dooley, F.L. Dryer, B. Yang, J. Wang, T.A. Cool, T. Kasper, N. Hansen, Combust.

Flame 158 (2011) 732.

[72] J.S. Francisco, J. Am. Chem. Soc. 125 (2003) 10475.

[73] W.K. Metcalfe, J.M. Simmie, H.J. Curran, J. Phys. Chem. A 114 (2010) 5478.

[74] S.L. Peukert, R. Sivaramakrishnan, M.-C. Su, J.V. Michael, Combust. Flame 159 (2012)

2312.

[75] C.K. Westbrook, W.J. Pitz, P.R. Westmoreland, F.L. Dryer, M. Chaos, P. Osswald, K.

Kohse-Höinghaus, T.A. Cool, J. Wang, B. Yang, N. Hansen, T. Kasper, Proc. Combust. Inst.

32 (2009) 221.

[76] R. Sumathi, J. William H. Green, Phys. Chem. Chem. Phys. 5 (2003) 3402.

[77] J.Y.W. Lai, K.C. Lin, A. Violi, Prog. Energy Combust. Sci. 37 (2011) 1.

[78] P.S. Nigam, A. Singh, Prog. Energy Combust. Sci. 37 (2011) 52.

[79] J.Q.A. Brito, C.S. Silva, J.S. Almeida, M.G.A. Korn, M. Korn, L.S.G. Teixeira, Fuel

Process. Technol. 95 (2012) 33.

[80] L. Coniglio, H. Bennadji, P.A. Glaude, O. Herbinet, F. Billaud, Prog. Energy Combust. Sci.

39 (2013) 340.

[81] B. Akih-Kumgeh, J.M. Bergthorson, Energy Fuels 25 (2011) 4345.

[82] W.K. Metcalfe, S. Dooley, H.J. Curran, J.M. Simmie, A.M. El-Nahas, M.V. Navarro, J.

Phys. Chem. A 111 (2007) 4001.

[83] W.K. Metcalfe, C. Togbé, P. Dagaut, H.J. Curran, J.M. Simmie, Combust. Flame 156

(2009) 250.

113

[84] S.M. Walton, M.S. Wooldridge, C.K. Westbrook, Proc. Combust. Inst. 32 (2009) 255.

[85] B. Yang, C.K. Westbrook, T.A. Cool, N. Hansen, K. Kohse-Höinghaus, Phys. Chem.

Chem. Phys. 13 (2011) 6901.

[86] G. Dayma, F. Halter, F. Foucher, C. Mounaim-Rousselle, P. Dagaut, Energy Fuels 26

(2012) 6669.

[87] W. Ren, J.B. Jeffries, R.K. Hanson, Meas. Sci. Technol. 21 (2010) 105603.

[88] E.R. Ritter, J.W. Bozzelli, Int. J. Chem. Kinet. 23 (1991) 767.

[89] C. Muller, V. Michel, G. Scacchi, G.M. Come, J Chim Phys 92 (1995) 1154.

[90] A.T. Blades, Can. J. Chem. 32 (1954) 366.

[91] J.-G. Chang, H.-T. Chen, S. Xu, M.C. Lin, J. Phys. Chem. A 111 (2007) 6789.

[92] N. Akiya, P.E. Savage, Aiche J. 44 (1998) 405.

[93] K. Saito, T. Shiose, O. Takahashi, Y. Hidaka, F. Aiba, K. Tabayashi, J. Phys. Chem. A 109

(2005) 5352.

[94] S.L. Fischer, F.L. Dryer, H.J. Curran, Int. J. Chem. Kinet. 32 (2000) 713.

[95] N. Leplat, J. Vandooren, Combust. Flame 159 (2012) 493.

[96] A. Galano, J.R. Alvarez-Idaboy, M.E. Ruiz-Santoyo, A. Vivier-Bunge, J. Phys. Chem. A

106 (2002) 9520.

[97] K. Saito, T. Sasaki, I. Yoshinobu, A. Imamura, Chem. Phys. Lett. 170 (1990) 385.

[98] J.C. Mackie, K.R. Doolan, Int. J. Chem. Kinet. 16 (1984) 525.

[99] A.V. Joshi, H. Wang, Int. J. Chem. Kinet. 38 (2006) 57.

[100] T.L. Nguyen, B.C. Xue, R.E. Weston, J.R. Barker, J.F. Stanton, J. Phys. Chem. Lett. 3

(2012) 1549.

[101] J.P. Senosiain, C.B. Musgrave, D.M. Golden, Int. J. Chem. Kinet. 35 (2003) 464.

[102] L.R. McCunn, K.-C. Lau, M.J. Krisch, L.J. Butler, J.-W. Tsung, J.J. Lin, J. Phys. Chem. A

110 (2006) 1625.

[103] L.K. Huynh, K.C. Lin, A. Violi, J. Phys. Chem. A 112 (2008) 13470.

[104] J.R. Barker, private communication (2013).

[105] D.R. Haylett, D.F. Davidson, R.K. Hanson, Combust. Flame 159 (2012) 552.

[106] M.F. Campbell, D.F. Davidson, R.K. Hanson, C.K. Westbrook, Proc. Combust. Inst. 34

(2013) 419.

[107] J. Herzler, M. Fikri, K. Hitzbleck, R. Starke, C. Schulz, P. Roth, G.T. Kalghatgi, Combust.

Flame 149 (2007) 25.

114

[108] Y. Sakai, A. Miyoshi, M. Koshi, W.J. Pitz, Proc. Combust. Inst. 32 (2009) 411.

[109] H. Wang, E. Dames, B. Sirjean, D.A. Sheen, R. Tangko, A. Violi, J.Y.W. Lai, F.N.

Egolfopoulos, D.F. Davidson, R.K. Hanson, C.T. Bowman, C.K. Law, W. Tsang, N.P.

Cernansky, D.L. Miller, R.P. Lindstedt, Jetsurf 20 (2010) http://melchior.usc.edu/JetSurF.

[110] S. Dooley, S.H. Won, M. Chaos, J. Heyne, Y. Ju, F.L. Dryer, K. Kumar, C.-J. Sung, H.

Wang, M.A. Oehlschlaeger, R.J. Santoro, T.A. Litzinger, Combust. Flame 157 (2010) 2333.

[111] E. Ranzi, A. Frassoldati, R. Grana, A. Cuoci, T. Faravelli, A.P. Kelley, C.K. Law, Prog.

Energy Combust. Sci. 38 (2012) 468.

[112] S. Gaïl, S.M. Sarathy, M.J. Thomson, P. Diévart, P. Dagaut, Combust. Flame 155 (2008)

635.

[113] H.J. Curran, P. Gaffuri, W.J. Pitz, C.K. Westbrook, Combust. Flame 129 (2002) 253.

[114] H. Wang, M. Frenklach, Combust. Flame 110 (1997) 173.

[115] S.W. Sharpe, T.J. Johnson, R.L. Sams, P.M. Chu, G.C. Rhoderick, P.A. Johnson, Appl.

Spectrosc. 58 (2004) 1452.

[116] L.T. Molina, W.B. Grant, Appl. Opt. 23 (1984) 3893.