You are on page 1of 120

Using an Opposed Flow Diffusion Flame to Study the Oxidation of C4 Fatty Acid Methyl Esters

by

Subram Maniam Sarathy

A thesis submitted in conformity with the requirements for the degree of Master of Applied Science Graduate Department of Chemical Engineering and Applied Chemistry University of Toronto

Copyright c 2006 by Subram Maniam Sarathy

Abstract
Using an Opposed Flow Diusion Flame to Study the Oxidation of C4 Fatty Acid Methyl Esters Subram Maniam Sarathy Master of Applied Science Graduate Department of Chemical Engineering and Applied Chemistry University of Toronto 2006 The oxidation of saturated (i.e methyl butanoate) and unsaturated (i.e. methyl crotonate) C4 fatty acid methyl esters in an opposed ow diusion ame has been studied to better understand the role of molecular structure in biodiesel combustion. The results indicate that the methyl crotonate ame produces higher levels of unsaturated hydrocarbons, which are known soot precursors in combustion applications. In addition, higher levels of acrolein, acetone, methanol, and benzene are observed in the methyl crotonate ame. The double bond in methyl crotonate is identied as the reason for the observed dierences in species formation. The experiments are also used to validate an improved detailed chemical kinetic model for methyl butanoate. The model exhibits good agreement with the experimental data. The major reaction pathways for methyl butanoate oxidation in the opposed ow diusion ame are presented herein.

ii

Sabbo pajjalito loko, sabbo loko pakampito The entire universe is nothing but combustion and vibration -Buddha-

Dedication
To all beings. May you be happy and peaceful. May you enjoy good health and harmony.

Acknowledgements
Deepest thanks to Professor Murray Thomson for his guidance and support in my academic endeavors. Je remercie Monsieur Sandro Ga Docteur en cintique chimique et combustion, pour l, e avoir partag ses connaissances en chimie de la combustion et sur les mthodes analye e tiques, et en plus, pour avoir attis mon intrt en franais. e ee c Warm thanks to Sajid Syed, my predecessor, who provided the experimental foundation and training I needed to be successful. My appreciation towards Professor Boocock and Professor Wallace for being members on my MASc defense committee. Thanks to AUTO21 and NSERC for providing research funding. My blessings to my twin brother, my wife, and all my family and friends. Words cannot express my gratitude towards you, so here is a loving smile. :)

iii

Contents
1 Introduction 1.1 1.2 1.3 Research Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Research Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Research Execution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1 2 2 4 4 5 6 7 7 10 11 13 14 14 14 15 16 16 17 19 20 20 21 22

2 Background Research 2.1 2.2 Basics of Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chemistry of Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 2.2.2 2.3 2.4 Feedstocks for Biodiesel . . . . . . . . . . . . . . . . . . . . . . . Surrogates of Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . Production of Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . Properties of Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . Biodiesel Use in Diesel Engines . . . . . . . . . . . . . . . . . . . . . . . 2.4.1 2.4.2 2.4.3 2.5 2.5.1 2.5.2 NOx Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . Biodiesel as a Lubricity Additive . . . . . . . . . . . . . . . . . . Oxygenated Fuels as Soot-Reducing Additives . . . . . . . . . . . Chemical Kinetic Modeling . . . . . . . . . . . . . . . . . . . . . Oxidation of Hydrocarbons . . . . . . . . . . . . . . . . . . . . . Intermediate and Low Temperature Oxidation of Hydrocarbons . High Temperature Oxidation of Alkanes . . . . . . . . . . . . . . Oxidation of Alkenes . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.3 Oxidation of Methyl Esters . . . . . . . . . . . . . . . . . . . . . Oxidation of Methyl Acetate . . . . . . . . . . . . . . . . . . . . . Oxidation of Dimethyl Carbonate . . . . . . . . . . . . . . . . . . Oxidation of Methyl Butanoate . . . . . . . . . . . . . . . . . . . iv

Modeling the Oxidation of Biodiesel . . . . . . . . . . . . . . . . . . . . .

2.6

Mechanism of Soot Formation in Combustion Processes . . . . . . . . . .

22 26 28 30 32 33 33 35 36 36 37 37 38 39 40 41 42 43 43 45 46 46 47 49

3 Experimental Apparatus and Analytical Methodology 3.1 3.2 3.3 3.4 Opposed ow diusion burner setup . . . . . . . . . . . . . . . . . . . . . Fuel Preparation and Vaporization . . . . . . . . . . . . . . . . . . . . . Supply of Fuel and Oxidizer Streams . . . . . . . . . . . . . . . . . . . . Gas Sampling System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.1 3.4.2 3.5 3.5.1 3.5.2 Sampling Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . Sampling Procedure . . . . . . . . . . . . . . . . . . . . . . . . . Non-Dispersive Infrared Analysis . . . . . . . . . . . . . . . . . . CO and CO2 Measurements . . . . . . . . . . . . . . . . . . . . . Gas Chromatography . . . . . . . . . . . . . . . . . . . . . . . . . GC Carrier Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . Injection System . . . . . . . . . . . . . . . . . . . . . . . . . . . GC Measurement Procedures . . . . . . . . . . . . . . . . . . . . Methodology for Analysis of Hydrocarbons . . . . . . . . . . . . . Methodology for Analysis of Oxygenates . . . . . . . . . . . . . . GC Calibration Procedure . . . . . . . . . . . . . . . . . . . . . . 3.5.3 High Pressure Liquid Chromatography . . . . . . . . . . . . . . . DNPH Sampling Procedure . . . . . . . . . . . . . . . . . . . . . HPLC Measurement Procedure . . . . . . . . . . . . . . . . . . . 3.5.4 Temperature Measurement . . . . . . . . . . . . . . . . . . . . . . Correction for Radiation Losses . . . . . . . . . . . . . . . . . . . 4 Modeling 4.1 4.2 Modeling Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Details of Input Files . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Analytical Tehniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

49 50 52 52 52 54 57

5 Results and Discussion 5.1 The Role of FAME Molecular Structure in Combustion . . . . . . . . . . 5.1.1 5.1.2 5.1.3 Major Reaction Pathways for the Oxidation of Methyl Butanoate Major Reaction Pathways for the Oxidation of Methyl Crotonate Comparison of Opposed Flow Diusion Flame Emissions Proles . v

5.1.4 5.2

Rationale for Dierences in Methyl Butanoate and Methyl Crotonate Emissions Proles . . . . . . . . . . . . . . . . . . . . . . . . 64 66 66 67 70 71 71 74 77 77 78 80 84 85 88 90 94

Chemical Kinetic Modeling of Methyl Butanoate . . . . . . . . . . . . . . 5.2.1 5.2.2 5.2.3 Modication of Reaction Rates - Contributions of this Study . . . Model Validation with Opposed Flow Diusion Flame Results . . Error Analysis of Major Dierences in Modeling and Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Experimental Errors . . . . . . . . . . . . . . . . . . . . . . . . . Inaccuracies in Model Reaction Rate Coecients . . . . . . . . . 5.2.4 Major Reaction Pathways for the Oxidation of Methyl Butanoate in the Opposed Flow Diusion Flame . . . . . . . . . . . . . . . .

6 Conclusions and Recommendations 6.1 6.2 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Bibliography Appendices A Flow Rates of Fuel and Oxidizer B Structures of Chemical Compounds C Sample Chromatograms D Sample Calculations

vi

List of Tables
2.1 2.2 2.3 2.4 Saturated fat composition of dierent vegetable oil and animal fats . . . Short chain and long chain fatty acid methyl esters . . . . . . . . . . . . Properties of biodiesels compared to standard diesel . . . . . . . . . . . . Relative magnitudes of rate constants for H abstraction from dierent CH bonds [38] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 3.2 3.3 3.4 3.5 Physical properties of the fuels used . . . . . . . . . . . . . . . . . . . . . Settings for the gas sampling valve and split valve . . . . . . . . . . . . . Flow rate of GC gases when measuring hydrocarbons . . . . . . . . . . . Flow rate of GC gases when measuring oxygenates . . . . . . . . . . . . . Determination of ideal sampling times for carbonyl compounds using the DNPH technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1 5.2 5.3 5.4 5.5 A comparison of measured peak species mole fractions in methyl butanoate (MB) and methyl crotonate (MC) ames . . . . . . . . . . . . . . . . . . Modied reaction rate constants [10, 8, 43]. . . . . . . . . . . . . . . . . . Decomposition pathways of CH3 OC =O and corresponding reaction rate constants [43]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A comparison of peak species mole fractions from experimental data and model-predicted values . . . . . . . . . . . . . . . . . . . . . . . . . . . . A comparison of reaction rates given in the present model [10] to those available in literature [67] . . . . . . . . . . . . . . . . . . . . . . . . . . B.1 Structures of relevant chemical species . . . . . . . . . . . . . . . . . . . 75 89 75 67 63 66 45 18 31 41 41 42 6 8 11

vii

List of Figures
2.1 2.2 2.3 2.4 2.5 2.6 2.7 Composition of a Triglyceride from [16] . . . . . . . . . . . . . . . . . . . Biodiesel Production and Blending Flowchart . . . . . . . . . . . . . . . Transesterication Reaction . . . . . . . . . . . . . . . . . . . . . . . . . Decay of isopropyl radical . . . . . . . . . . . . . . . . . . . . . . . . . . Addition of O radical to ethene from [38] . . . . . . . . . . . . . . . . . . Structure of Methyl Acetate . . . . . . . . . . . . . . . . . . . . . . . . . Main reaction pathways for DMC in the opposed ow diusion ame (percentages refer to a reaction pathways share of a species consumption) from [43] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.8 2.9 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 Main reaction pathways for the oxidation of methyl butanoate from Fisher et al. [8] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . General mechanism for soot formation from Glassman [38] . . . . . . . . Schematic of the experimental setup . . . . . . . . . . . . . . . . . . . . 23 25 27 29 30 32 34 35 37 38 39 40 40 42 21 5 8 9 19 19 20

Diagram of burner port . . . . . . . . . . . . . . . . . . . . . . . . . . . . Photograph of burner setup . . . . . . . . . . . . . . . . . . . . . . . . . Photograph of fuel vaporization column . . . . . . . . . . . . . . . . . . . Schematic of microprobe (not to scale) . . . . . . . . . . . . . . . . . . . Schematic of microprobe and burner setup (not to scale) . . . . . . . . . Schematic of NDIR instrument from [55] . . . . . . . . . . . . . . . . . . Schematic of GC instrument . . . . . . . . . . . . . . . . . . . . . . . . . Schematic of FID from [57] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3.10 Schematic of injector in the ll position . . . . . . . . . . . . . . . . . . . 3.11 Schematic of injector in the load position 3.12 Oven temperature program for measuring hydrocarbons using HP Al/S PLOT column . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

3.13 Oven temperature program for DB-624 Column . . . . . . . . . . . . . . 3.14 Schematic of HPLC Setup from [62] . . . . . . . . . . . . . . . . . . . . . 3.15 Thermocouple Schematic . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1 5.1 5.2 5.3 5.4 5.5 5.6 5.7 Geometry of opposed-ow diusion ame from [64] . . . . . . . . . . . . Hydrogen abstraction from carbons atoms in methyl butanoate . . . . . . Decomposition reaction pathways for methyl butanoate . . . . . . . . . . Reaction pathway for forming methyl crotonate from methyl butanoate . Hydrogen abstraction from carbons atoms in methyl crotonate . . . . . . Decomposition reaction pathways for methyl crotonate . . . . . . . . . . Reactions with biradical O for methyl crotonate . . . . . . . . . . . . . . Diagram of the burner setup to clarify the orientation of experimental proles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.8 Measured temperature proles in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.9 Measured concentration proles for fuel (MB or MC), CO, and CO2 in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames . . . . . . . . . . . . . . . . . 5.10 Measured concentration proles for CH4 , C2 H4 , and C2 H2 in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC open symbols without lines) ames . . . . . . . . . . . . . . . . . . . . . 5.11 Measured concentration proles of C3 H6 and C2 H6 in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames, and CH3 OH in the methyl crotonate ame . . . . . 5.12 Measured concentration proles of 2-C4 H6 , C3 H8 , and CH3 CHO in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames . . . . . . . . . . . . . . . . . 5.13 Measured concentration proles of C4 H8 , C3 H4 , 1,3-C4 H6 , and CH2 O in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames . . . . . . . . . . . . . . . ix

43 44 47 50 53 55 55 56 56 58

58

59

59

60

60

61

61

5.14 Measured concentration proles of C5 H12 and C4 H10 in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC open symbols without lines) ames, and C6 H6 , C3 H4 O, and C3 H6 O in the methyl crotonate ame . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.15 Analogous reaction pathways for methyl butanoate and methyl crotonate oxidation in the opposed ow diusion ame . . . . . . . . . . . . . . . . 5.16 Modeling predictions (small symbols with lines) and experimental results (large symbols without lines) for methyl butanoate (MB), CO, and CO2 . 5.17 Modeling predictions (small symbols with lines) and experimental results (large symbols without lines) for CH4 , C2 H4 , and C2 H2 . . . . . . . . . . 5.18 Modeling predictions (small symbols with lines) and experimental results (large symbols without lines) for C3 H6 , C2 H6 , and CH2 O . . . . . . . . . 5.19 Modeling predictions (small symbols with lines) and experimental results (large symbols without lines) for C4 H8 , C3 H8 , pC3 H4 , CH3 CHO, and 1,3C 4 H6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.20 Primary reaction pathways for methyl butanoate oxidation in the opposed ow diusion ame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.1 Example of HPLC chromatogram of DNPH derivatives . . . . . . . . . . C.2 Example of GC chromatogram for the HP Plot Column . . . . . . . . . . C.3 Example of GC chromatogram for the DB 624 Column . . . . . . . . . . 76 91 92 93 69 69 68 68 65 62

Chapter 1 Introduction
There is a need for renewable fuels in North America due to the negative impacts of nonrenewable fuels (petroleum) on the environment and societies. Among renewable fuels, those derived from biomass sources (biofuels) are gaining importance. The primary drivers for using biofuels are to reduce greenhouse gas emissions and consumption of fossil fuels. Reducing harmful emissions to the atmosphere is imperative for mitigating global warming and sustaining healthy metropolitan areas for human inhabitation. In addition, the ability for citizens to locally grow and produce their own fuel minimizes the dependence on nonrenewable energy sources. The Canadian Greenhouse Gas Inventory [1] reported that the transportation sector accounts for nearly 25% of all greenhouse gas emissions to the atmosphere. The majority of these emissions came from gasoline vehicles (53%) and diesel vehicles (35%) [1]. Both the gasoline engine and the diesel engine are capable of operating on biofuels. The respective biofuels of interest for gasoline and diesel technologies are ethanol and biodiesel. In North America, ethanol has been blended with gasoline for over a decade; however, the use of biodiesel has only started gaining momentum in the past few years. Rudolf Diesel, the inventor of the diesel engine stated in 1912 that, the use of vegetable oils for engine fuels may seem insignicant today. But such oils may become in course of time as important as petroleum and the coal tar products of the present time [2]. Biodiesel is dened as a mixture of mono-alkyl esters of long chain fatty acids derived from vegetable oils or animal fats [3]. Biodiesel can be used in its pure form or it can be blended with petroleum diesel without major modications to the existing engine and fuel distribution infrastructure [4]. Engine combustion studies [5, 6, 7] have shown that the use of biodiesel can signicantly reduce hydrocarbon, particulate, carbon monoxide, 1

Chapter 1. Introduction

and greenhouse gas emissions while increasing fuel economy; however, a slight increase in NOx emissions have been observed. These benets can be optimized by understanding the fundamentals of biodiesel combustion.

1.1

Research Motivation

In order to design engine systems, designers require information about the fundamental combustion properties of the fuel. However, detailed studies have not been conducted for biodiesel. In addition, the properties of each biodiesel variant depend on the feedstock used for its preparation, i.e., biodiesel derived from animal fats performs dierently from biodiesel derived from vegetable oils. This lack of information has inhibited the widespread use of biodiesel in the automotive sector because engine manufacturers are reluctant to design systems for a poorly understood fuel. Detailed chemical kinetic models are an essential input into the engine computational uid dynamic (CFD) models used by designers. There are currently no validated chemical kinetic models for the long chain fatty acid methyl esters (FAME) found in biodiesel. However, Fisher et al. [8] have provided a chemical kinetic mechanism for methyl butanoate. Methyl butanoate is a fully saturated short chain FAME with a molecular structure similar to that of the saturated long chain FAMEs found in biodiesel. It was chosen as a modeling surrogate of biodiesel because it is large enough to allow fast RO2 isomerization reactions important in the low-temperature chemistry that controls fuels auto-ignition under conditions found in diesel engines [8]. The authors were not able to robustly validate the model due to limited experimental data on methyl butanoate combustion. Previously, Syed [9] provided opposed ow diusion ame data to validate the Fisher et al. model [8]; however, measurements for higher molecular weight hydrocarbons and important oxygenated species were not provided. Following the Syed study [9], an improved methyl butanoate oxidation model has been developed by Dr. Sandro Ga and l coworkers [10]. However, the existing opposed ow diusion ame data is not sucient to validate the Ga et al. model [10]. l In addition, methyl butanoate is not a good surrogate fuel for the unsaturated FAMEs found in some biodiesels. It has been shown [6, 7, 11] that biodiesel combustion performance depends on the molecular structure of the fuel (i.e. extent of unsaturation). Therefore, it would be interesting to understand the role of molecular structure during

Chapter 1. Introduction

combustion by comparing ame emissions and oxidation reaction pathways for saturated and unsaturated FAMEs.

1.2

Research Objectives

The ultimate goal of our research is to develop detailed chemical kinetic models for the oxidation of biodiesel, and then validate the models using a number of experimental platforms. However, for this study the scope has been reduced to the oxidation of C4 fatty acid methyl esters in an opposed ow diusion ame. The goal to oxidize methyl butanoate (saturated) and methyl crotonate (unsaturated) in an opposed ow diusion ame, and then measure the ame species proles to: 1. determine the relationship between the FAME molecular structure (i.e. unsaturation) on the ame characteristics; 2. and validate an existing chemical kinetic model for methyl butanoate.

1.3

Research Execution

Laboratory experiments were conducted at the Combustion Research Laboratory1 . Flame studies were carried out at atmospheric pressure in an opposed ow diusion ame, which burns the fuel at a stagnation point, so only combustion kinetics, diusion (i.e. no turbulent mixing), and thermodynamics inuences the emissions formed. Methyl butanoate (saturated) and methyl crotonate (unsaturated) were the two fuels tested. The emissions were sampled by drawing the gases from within the ame using a quartz microprobe. The emissions were characterized using a number of analytical techniques. CO and CO2 concentrations were measured using nondispersive infrared (NDIR) spectroscopy. Hydrocarbon concentrations were obtained by gas chromatography with a ame ionization detector (GC/FID). High pressure liquid chromatography (HPLC) was used to measure concentrations of formaldehyde and other oxygenated species. The structure of the opposed ow diusion ame was modeled using a modied version of the one-dimensional steady-state ame code OPPDIF and the chemical kinetic model for methyl butanoate oxidation by Dr. Ga [10]. The chemical kinetic mechanism was l
1

University of Toronto Department of Mechanical and Industrial Engineering

Chapter 1. Introduction

then validated against the experimental data. After validation, the model was used to analyze the major reaction pathways for methyl butanoate combustion. Pathway analysis was further conducted to explain the role of molecular structure in combustion.

Chapter 2 Background Research


2.1 Basics of Biodiesel

ASTM International denes biodiesel as a fuel comprised of mono-alkyl esters of long chain fatty acids derived from vegetable oils or animal fats, designated B100, and meeting the requirements of ASTM D6751 [3]. Biodiesel refers to the pure fuel before blending with diesel fuel. Biodiesel blends are denoted as, BXX with XX representing the percentage of biodiesel contained in the blend (ie: B20 is 20% biodiesel, 80% petroleum diesel) [4]. Biodiesel oers many advantages [12]: It is a renewable fuel. It displaces the use of petroleum diesel. It is less toxic than table salt and more biodegradable than white sugar. It can help reduce greenhouse gas emissions. It can reduce particulate matter (PM), carbon monoxide (CO), sulphur oxides (SOx ), and hydrocarbon (HC) emissions. It can be made locally from agricultural and/or recycled feedstocks. Biodiesel can be used in its pure form (i.e. B100), or it can be blended with petroleum diesel, as is more common in practice. Suppliers can blend 1-2% biodiesel as a lubricity additive, which will become more important when ultra low sulfur diesel fuels (ULSD) 5

Chapter 2. Background Research

are mandated in June, 2006. A higher blend of 20% biodiesel (i.e. B20) is the most common in the United States because it balances performance, cost, emissions, and handling with petroleum diesel. Higher blend levels are currently not recommended because modications and special care may be required depending on the type of engine and vehicle being used [12]. It should be noted that raw or rened fats and oils that have not been processed to meet the ASTM D6751 standards for use in diesel engines are not considered biodiesel and should be avoided. The main purpose of processing the oil to biodiesel is to change the fuels viscosity. Diesel engines, and in particular the fuel injectors, are designed for diesel fuel with a viscosity of between 1.3-4.1 mm2 /s; however, raw oils tend to be around 40 mm2 /s [12]. Research [13, 14, 15] has shown that blends using raw oils and greases can cause severe long-term damage to the engine system. Therefore, oils must be processed into biodiesel to reduce the fuels viscosity to a value similar to petroleum diesel, usually 4-5 mm2 /s [12].

2.2

Chemistry of Biodiesel

As previously mentioned, biodiesel is derived from fats and oils. Oils and fats are made of compounds termed triglycerides. Triglycerides consist of one molecule of glycerol combined with three molecules of fatty acid, as shown in Figure 2.1. A fatty acid is a hydrocarbon chain terminating in a carboxyl group, as shown in Equation 2.1. If the three fatty acids in the triglyceride are identical, then it is a simple triglyceride. However, most triglycerides contain a mixture of dierent fatty acids. Thus, the properties of these mixed triglycerides depend on the properties of the fatty acids present. Subsequently, since fats and oils are typically a mixture of dierent mixed triglycerides, their properties are ultimately determined by the fatty acid content.

H3 C CH2 .............CH2 CH2 COOH

(2.1)

The most important vegetable oils consist of fatty acids containing 16 and 18 carbon atoms [17]. If a double bond exists between two carbon atoms, then the chain is described as unsaturated because all available carbon valences for hydrogen are not satised. Palmitic acid, C15 H35 COOH, and stearic acid, C15 H35 COOH, are examples of saturated fatty acids, while oleic acid, CH3 (CH2 )7 CH=CH(CH2 )7 COOH, is an unsat-

Chapter 2. Background Research

Figure 2.1: Composition of a Triglyceride from [16]

urated fatty acid. Unsaturated fatty acids are associated with a lower melting condition, greater solubility and chemical reactivity. In contrast, saturated fatty acids have higher melting characteristics, and are less reactive and soluble. Unsaturated fatty acids with two or more double bonds are called polyunsaturated fatty acids (PUFAs). The extent of unsaturation is quantied by the iodine value (IV). It is determined by the amount of iodine taken up by 100 grams of a fatty acid. Saturated fatty acids take up no iodine; therefore, their IV is zero [18]. In general, vegetable oils tend to contain more unsaturated fatty acids while animal fats contain more saturated fatty acids. However, this is not always true as shown in Table 2.1. It is clear that palm oil and coconut oil do not follow the general properties of most vegetable oils [19].

2.2.1

Feedstocks for Biodiesel

There are a variety of feedstocks available for biodiesel production, but most can be classied as vegetable oils, rst-use animal fats, or waste greases [19]. The physical chemical properties and molecular structure of a given biodiesel are directly related to

Chapter 2. Background Research


Oil or Fat Canola Safflower Sunflower Corn Olive Soybean Chicken Fat Beef Tallow Palm Coconut Percent of Saturated Fats 7 9 10 13 13 17 30 47 50 97

Table 2.1: Saturated fat composition of dierent vegetable oil and animal fats

the properties of the fat or oil from which it was derived. Biodiesel derived from highly unsaturated feedstocks, such as canola and soybean oil, will be equally unsaturated. Similarly, biodiesel derived from animal fats, such as beef tallow, are highly saturated because the feedstock consists primarily of saturated fatty acids. Vegetable oils are virgin oils derived from any vegetable source, such as soybeans, canola, sunowers, peanuts, etc. Generally they are low in saturated fats, and therefore the least expensive to process. On the other hand, they are the most expensive feedstock to purchase. The lack of saturated esters produced from vegetable oils results in a biodiesel with better cold ow characteristic and lower cetane numbers. The presence of double bonds in the FAME also makes the fuel more susceptible to oxidation, thereby reducing its shelf-life [12]. First-use animal fats include any virgin fat or oil derived from an animal source, such as edible and inedible tallows, grades of lards, poultry fats, etc. These feedstocks are generally less expensive to purchase than vegetable oils. Biodiesel produced from rst-use animals fats are high in saturated fats; therefore, they are more expensive to produce, have poorer cold ow characteristics, and higher cetane numbers. Waste greases include waste vegetable oils (yellow greases) that were used for cooking and waste trap greases (brown greases). These are the least expensive feedstocks for biodiesel production, but they require the most intensive processing. Waste greases

Chapter 2. Background Research display characteristics of both vegetable oils and animal fats [20]

In addition to saturation, the free fatty acid (FFA) content in the feedstock oil or fat is also an important characteristic. Fatty acid chains become FFAs when they break o the triglyceride. These FFAs can aect the biodiesel production process, as will be discussed in Section 2.2.2 . The FFA content of various biodiesel feedstocks is as follows [20]: Rened vegetable oils, such as soybean or canola oil (FFA < 1.5%) Low FFA yellow greases (waste vegetable oils)1 and animal fats (FFA < 4%) High FFA greases and animal fats (FFA 20%) Surrogates of Biodiesel Surrogates for biodiesel fuel molecules were used in this study to model the combustion characteristics of real biodiesel. The purpose of using surrogates is to simplify the combustion model by reducing the number of possible chemical reactions, while still representing the role of the molecular structure in combustion, i.e. the role of the methyl ester group and the role of carbon-carbon double bonds. In addition, real biodiesel is a complex mixture of fatty acid esters with diering chain lengths and degrees of unsaturation, thereby increasing the complexity of the modeling procedure. In this study, two short chain fatty acid methyl esters, methyl butanoate and methyl crotonate, were used as surrogates for the saturated and unsaturated long chain fatty acid methyl esters, respectively. The chemical formula and molecular weight of these compounds, along with similar long chain fatty acid methyl esters, is show in Table 2.2.

2.2.2

Production of Biodiesel

Producing biodiesel that meets the ASTM D6751 standards requires the feedstock oil or fat to be processed, so that its viscosity is reduced. Several methods by which the viscosity can be reduced are micro-emulsication, cracking, and transesterication. Ma and Hanna have reviewed these processes [21]; however, only transesterication produces a fuel that meets the ASTMs denition of biodiesel [3], wherein biodiesel is comprised
1

Yellow grease is another term for waste vegetable oils usually from restaurants

Chapter 2. Background Research

10

Table 2.2: Short chain and long chain fatty acid methyl esters

of long chain mono alkyl esters. A simplied owchart for biodiesel production and blending is shown in Figure 2.2.

Figure 2.2: Biodiesel Production and Blending Flowchart

Transesterication is the process by which a triglyceride (the basic constituent of fats and oils) is converted to a mono alkyl ester. It involves reacting the fat or oil with an alcohol, in the presence of a catalyst, to produce glycerol and esters [22]. Methanol is the most commonly used alcohol, therefore the resulting biodiesel is often referred to as a mixture of fatty acid methyl esters (FAMEs). From here on, FAME will be used to refer to pure fatty acid methyl ester compounds while biodiesel will be used to refer to a mixture of various fatty acid methyl esters. Ethanol has also been used [23], and in

Chapter 2. Background Research

11

such cases the resulting biodiesel is called fatty acid ethyl ester. The transesterication reaction between a typical triglyceride and methanol is shown in Figure 2.3

Figure 2.3: Transesterication Reaction

Three moles of methanol are required to react with each mole of triglyceride in the oil or fat. The reactions occur sequentially, i.e., the triglyceride reacts to form one mole of diglyceride and one mole of FAME, followed by the diglyceride reacting to one mole of monoglyceride plus one mole of FAME, and nally the monoglyceride to a mole of glycerol and the third mole of FAME [9]. Excess methanol is used to drive the reaction forward. The reaction is carried out in the presence of a catalyst to improve the reaction rate. Ramadhas et al. stated, enzymes, alkalis, or acids can catalyze the reaction, i.e. lipases, NaOH, and sulphuric acid, respectively [22]. Alkali catalysts such as sodium methoxide, sodium hydroxide, and potassium hydroxide are the most commonly used catalysts. Among these, sodium methoxide accounts for 70 percent of the biodiesel produced in North America [24]. A study by Vicente et al. at the University of Madrid [25] indicates that the highest yield of biodiesel (wt%) is obtained by sodium methoxide (99.33%), when compared to sodium hydroxide (86.71%) and potassium hydroxide (91.67%). However, sodium methoxide catalysts are the most expensive, and sodium hydroxide catalysts proved to be the fastest. Currently, only sodium methoxide is used in commercial applications. The catalyst, alcohol, and vegetable oil are combined in a batch reactor at approx-

Chapter 2. Background Research

12

imately 65 C and stirred continuously. The duration of the reaction can range from 1-6 hours, depending on the desired yield. Two immiscible layers are formed once the reversible reaction reaches equilibrium. The lower layer is glycerol and the upper layer contains FAMEs and unreacted feedstock. Many processes separate the glycerol and conduct a second transesterication reaction to increase the yield. Glycerol is a valuable by-product that can be sold for a prot. If the biodiesel is to meet ASTM D6751 specications, it must undergo a series of separation processes for the removal of alcohol, catalyst, water, soaps, glycerol, and unreacted triglycerides and free fatty acids (FFAs) [20]. It was previously mentioned that some feedstocks contain FFAs. FFAs react with the alkali catalyst to form soap and water during the transesterication process. The loss of catalyst can be compensated for by supplying additional catalyst. However, this is an expensive option and it does not address the issue of excess water, which can limit the transesterication. An expensive yet commonly used alternative is to pre-process the feedstock to remove the small quantities of FFAs. It is not economical to remove FFAs from high FFA greases and animal fats, and in such cases an acid-catalysis is rst used to convert FFAs to methyl esters, followed by standard base-catalyzed transesterication [20].

2.3

Properties of Biodiesel

Biodiesel must have physical-chemical properties similar to that of diesel fuel in order to successfully operate in a compression ignition engines. The properties which most commonly aect operation are viscosity, heating value, ignition quality (cetane number), pour point, and cloud point. This section summarizes a study by the NREL [20] on the important properties of several biodiesel fuels and standard low sulfur diesel, as quantied by various ASTM standard testing methods. It should be noted that all engine manufactures design their commercial diesel engine systems to operate of fuel that meets the ASTM D 975 standards - Standard Specication for Diesel Fuel Oils [26]. The heating value, or heat of combustion, is a measure of the fuels energy density. Diesel engines are capable of accepting a variation in heating values, so there is no specied heating value in the ASTM D 975 Standard. However, it is benecial for biodiesel to have a heating value similar to that of standard diesel [22]. Table 2.3 indicates that the heating value of biodiesel is slightly lower than that of standard diesel. Therefore,

Chapter 2. Background Research on a weight basis, biodiesel is slightly less energy ecient than standard diesel.

13

The cetane number is a measure of the fuels ignition delay. Diesel combustion requires the fuel to self-ignite as it is sprayed into the compressed cylinder gas. The self-ignition leads to the characteristic diesel knock, wherein an explosion of premixed air and fuel causes a rapid heat release and a rapid pressure rise. The magnitude of the explosion can be decreased but shortening the ignition delay time. Higher cetane numbers result in shorter ignition delay times, and therefore better engine operation. The ASTM D 975 minimum acceptable cetane value is 40. Table 2.3 indicates that all biodiesels have higher cetane numbers when compared to standard diesel. Therefore, biodiesel is expected to improve engine operation. The viscosity measures the fuels resistance to ow. Diesel engine fuel injectors are highly sensitive to fuel viscosity. Higher viscosity results in poorer atomization of the fuel spray and an increase in engine deposits. The ASTM D 975 maximum acceptable value for viscosity is 4.1 mm2 /sec. Table 2.3 indicates that all biodiesels have a viscosity over two times greater than the viscosity of conventional diesel. Due to such high viscosities, pure biodiesel (B100) is not certied for use in most diesel engines. The fuel viscosity can be reduced by blending the biodiesel with conventional diesel fuel or another viscosity reducing additive. The pour point is a measure of the lowest possible temperature at which the fuel is observed to ow. At temperatures below the pour point, the fuel becomes gelled and can no longer ow. It is important that the pour point temperature be as low as possible, so that the fuel can be used in colder climates without the use of heaters. Table 2.3 shows the pour point temperatures of various fuels. In general, all biodiesels have signicantly higher pour points than conventional diesel. Biodiesel derived from animal fats and yellow greases have higher pour points due to the higher proportion of saturated FAMEs. The cloud point measures the temperature at which the fuel becomes hazy due to crystallization of the fuel particles. Using a product below its cloud point can plug certain lters in the fuel delivery system. Table 2.3 shows that the cloud point of all biodiesels is much higher than standard diesel. As expected, biodiesels with more saturated FAMEs have higher cloud point temperatures. The use of additives can help to improve the cold ow properties of biodiesel. ASTM D 975 does not specify a single number for cloud point and pour point temperatures. Instead, values are specied depending on the typical ambient temperature in the region of use. The specications indicate that 41 of the 48 states in the conti-

Chapter 2. Background Research

14

nental US reach temperatures below the lowest cloud point and pour point temperatures of biodiesel. Therefore, the use of biodiesel in its pure form (B100) is not possible in most regions; however, property-enhancing additives and petroleum diesel blending can be used to improve the cold temperature performance of biodiesel.
Physical Chemical Property Low FFA Yellow Grease ME High FFA Yellow Grease ME 2D Diesel ASTM D975 Reference Specication Fuel

Soy ME

Canola ME

Lard ME

Edible Tallow ME

Inedible Tallow ME

Heating Value (Btu/lb)

17153

17241

17165

17144

17061

17215

17154

18600

N/A

Cetane Number

59

53.9

N/A

74.8

54.3

55.2

53.2

47

> 40

Viscosity (mm2/sec)

4.55

4.63

4.85

4.91

4.93

5.62

4.66

2.45

1.9 - 4.1

Cloud Point (C)

-3

14

20

23

42

-18

by customer

Pour Point ( C)

-1

-4

11

13

12

-27

by customer

Table 2.3: Properties of biodiesels compared to standard diesel

2.4

Biodiesel Use in Diesel Engines

There have been a number of studies on the use of biodiesel in diesel engines. The goal of these studies was to determine the eect of using biodiesel and diesel-biodiesel blends on the emissions, fuel economy, and operation of the diesel engine. In addition, several experiments were performed to study the eects of biodiesel feedstock on the resultant fuels combustion properties. Following is a summary of various studies. Ramadhas et al. [22] conducted an extensive literature review on previous biodiesel engine studies. The wide variation in some data led them to some rather general conclusions below: Cold weather operation with biodiesel is not easy.

Chapter 2. Background Research

15

Additives are needed to improve the cold ow properties, material compatibility, oxidation stability, and viscosity. The use of biodiesel results in a thermal eciency comparable to conventional diesel. Small amounts of power loss (brake horse power) can be expected when using biodiesel. Biodiesel gives performance and emissions characteristics similar to those of conventional diesel, and biodiesel is an appropriate substitute for diesel fuel. Agarwal and Das [27] conducted performance and emission tests to evaluate performance of linseed oil methyl ester blends in diesel engines. Their study concluded, that the biodiesel can be used as an alternative, environment friendly fuel in existing diesel engines without substantial hardware modication. In addition, the optimal concentration of biodiesel in the blend was found to be 20%. It gave net advantage of 2.5 percent in peak thermal eciency and there was substantial reduction in smoke opacity values. However, the 20% biodiesel blend led to a 5% increase in NOx emissions. NOx is a highly temperature dependent phenomenon, and it was found that increasing the concentration of biodiesel led to a increase in exhaust gas temperature and NOx emissions. The United States Environmental Protection Agency [7] conducted an analysis on pre-existing biodiesel emissions data to compare emissions proles for three B20 blends against standard diesel in heavy-duty applications. The report indicates that a soybeanbased B20 reduces average emissions of PM by 10%, HCs by 20%, CO by 11%, and increases NOx emissions by 2% and fuel consumption by 1-2%. Rapeseed-based B20 produced similar results. However, it was found that animal fat-based biodiesel emits lesser quantities of NOx , PM and CO compared to vegetable oil-based fuels. Durbin and coworkers [6] studied the exhaust emissions of various B20 (2 soybean based and 1 yellow grease based) relative to standard diesel. Experiments were performed by chassis dynamometer tests on 7 light-duty diesel vehicles. The yellow grease biodiesel blend showed signicant reductions in HC and CO emissions across the test vehicle eet, with HCs reduced by 21-66% and CO reduced by 0-46%, relative to the standard diesel. HC and CO emissions were comparable for soybean methyl esters and standard diesel. The PM emissions for yellow grease biodiesel were only lower for the highest emitting vehicle, and comparable on the remaining vehicles. Soybean B20 had higher PM emission

Chapter 2. Background Research

16

rates than standard diesel for 4 of the 7 vehicles tested, and comparable PM emissions on the remaining vehicles. NOx emissions were not signicantly dierent between the any of the biodiesel blends and standard diesel. No correlation between FAME molecular structure and PM and NOx emissions could be drawn from this study. McCormick, Graboski and coworkers [11, 28, 29] have conducted many engine studies with biodiesel blends. Previous studies [28] showed that a 35% soybean methyl ester blend reduces PM emissions by 26% and increase NOx emissions by 1%, relative to standard diesel. B100 was found to increases NOx by 11% and reduce PM by 66%. Their most recent biodiesel study [11] tested biodiesels from various feedstocks in a heavy-duty engine to determine the impact of biodiesel chemical structure, i.e. fatty acid chain length and degree of unsaturation, on NOx and PM emissions. They correlated fuel density and cetane numbers with NOx and PM emissions: an increase in density correlates with an increase in NOx and little change in PM; an increase in cetane number correlates to a decrease in NOx and little change in PM. Saturated esters have higher cetane numbers and lower densities than unsaturated esters. Thus, NOx emissions are shown to increase with increasing iodine value. Decreasing fatty acid chain length was also shown to decrease NOx emissions. PM emissions were found to be unrelated to FAME molecular structure. Thus, they concluded that modifying the fuels molecular structure can alter biodiesel NOx emissions performance.

2.4.1

NOx Formation

Explanations for the observed increase in NOX in biodiesel are open for speculation. The three prevailing mechanisms explaining NOx formation in combustion engines are: 1.) Fuel, 2.) Thermal (Zeldovich mechanism), and 3.) Prompt (Fenimore mechanism). Fuel NOx is formed when nitrogen compounds xated in the fuel are oxidized. This is not a concern for biodiesel since the fuel does not contain any chemically bound nitrogen. Thermal NOx is a result of high temperature dissociation and chain reaction of elemental nitrogen and oxygen in the post-combustion regime. The reactions in this mechanism are shown in Equations 2.2, 2.3, and 2.4. The reactions are highly temperature dependent, such that a decrease in combustion temperature will decrease NOx formation [30]. However, a corresponding increase in PM will occur; thus the so-called NOx /PM tradeo. Agarwal and Das [27] provide engine studies supporting the thermal NOx theory. Tat and Van Gerpen [31] have shown that biodiesel has a higher isentropic bulk modulus

Chapter 2. Background Research

17

of compressibility than conventional diesel, which causes an inadvertent advance in fuel injection time in a pump-line-nozzle fuel injector. The advance in fuel injection timing increases the ignition delay, and therefore increases thermal NOx formation. Boehman and coworkers [32] have shown that the bulk modulus of compressibility increases with increasing iodine value. This explains why fuels with higher iodine values produce more NOx . However, engine studies by McCormick and coworkers [11] have shown that NOx emissions increase in unsaturated FAMEs while PM emissions remain unchanged. Therefore, the increased formation of NOx in biodiesel combustion cannot be entirely attributed to thermal NOx . This leads to the possibility that NOx is formed by the prompt NOx mechanism. Prompt NOx forms from a complex set of reactions between elemental nitrogen, hydrocarbon radicals, and oxygen under fuel-rich conditions [30]. It is proposed that combustion of unsaturated FAMEs leads to an increase of certain hydrocarbon radicals in the fuel rich zone of the diesel spray, and results in prompt NOx formation. Hess and coworkers [30] have shown that the use of antioxidants can terminate radical reactions and lead to reduction in NOx emissions. O + N2 N O + N O2 N + O2 N O + O N + HO N O + H (2.2) (2.3) (2.4)

2.4.2

Biodiesel as a Lubricity Additive

Conventional diesel fuel contains small amounts of sulphur containing molecules which serves as a lubricity agent, but at the same time increases SOx emissions. The mandate for ultra low sulfur diesel (ULSD) poses a concern since a reduction in diesel fuel lubricity can be damaging to the engine and fuel injection system. Biodiesel has been explored as an ULSD lubricity additive for extending the life of diesel engine components. The NREL [20] studied the properties, including lubricity, of biodiesel derived from various feedstocks. They reported that even small additions of biodiesel to diesel can signicantly improve lubricity. Lubricity was shown to improve by 10% with 0.25% biodiesel present, and by 30% with 0.5% biodiesel. Biodiesel derived from animal fat and yellow grease were found to increase lubricity better than biodiesel derived from vegetable

Chapter 2. Background Research

18

oils. However, the decrease in lubricity found in vegetable oils cannot be attributed to lower saturated FAME content, as is shown in the following study. Geller and Goodrum [33] examined the eect of FAME chain length and saturation on the fuels lubricity properties. They chose to study individual FAMEs rather than actual biodiesel fuels, which contain a mixture of FAMEs. No correlation between FAME chain length and lubricity enhancing properties was reported. However, saturated FAMEs were found to be less eective as a lubricity enhancer than unsaturated FAMEs. It was also found that hydroxylated FAMEs are the best at increasing lubricity. The author of this thesis hypothesizes that higher concentrations of hydroxylated FAMEs may explain the increased lubricity of biodiesel derived from animal fat, as seen in the aforementioned NREL study [20]. Hydroxylated FAMEs are usually derived from hydroxylated fatty acids which are found in small quantities in some oils and fats. Another possibility is that FFAs are converted to hydroxylated FAMEs during the transesterication process. Thus, the higher FFA content of animal fat and yellow grease would result in the formation of greater amounts of hydroxylated FAMEs. However, experimental data is not available to test this hypothesis.

2.4.3

Oxygenated Fuels as Soot-Reducing Additives

The use of oxygenated fuels has been shown to be an eective way of reducing soot emissions in diesel engines [34]. Oxygenated fuels prevent carbon atoms from forming soot by creating carbon-oxygen bonds. Ideally, each fuel bound O atom should pair up with a C atom to form CO. This prevents the formation of C-C bonds, which ultimately lead to soot formation. In addition, the oxidation of oxygenated species forms OH radicals, which readily attack unsaturated hydrocarbons and prevent their participation in soot growth reactions. It is for these reasons that the oxygenated methyl ester portion of FAMEs has made biodiesel an attractive blending agent for reducing diesel soot emissions. Several insights into the use of biodiesel for this purpose are presented in this study.

2.5

Modeling the Oxidation of Biodiesel

The combustion of fuel in the diesel engine is an oxidation reaction. Fuel is combined with hot compressed oxygen (air) at a ratio such that the fuel oxidizes rapidly. Chemical kinetic phenomena play an important role in understanding the initiation of the oxidation

Chapter 2. Background Research

19

process and the emissions formed during, and after, the combustion. There have not been any detailed studies on the chemical kinetic pathways for the oxidation of long chain FAMEs. However, numerous studies have been conducted for hydrocarbon fuels and oxygenated compounds. The development of chemical kinetic models for the oxidation of the FAMEs found in biodiesel can greatly benet from existing oxidation pathway studies for aliphatic hydrocarbons. FAMEs consist of a long hydrocarbon chain terminated by a carboxylate ester group. Therefore, it is likely that the oxidation of these compounds resemble the oxidation of alkanes and alkenes.

2.5.1

Chemical Kinetic Modeling

Detailed chemical kinetic mechanisms are used to model the molecular level transformation of reactants into products [35]. Take for example the oxidation of methane, as shown in Equation 2.5. Although the reaction looks simple, it actually proceeds through a number of elementary reactions. The rates of formation and consumption in the elementary reactions are described by a set of dierential equations, which are numerically integrated to determine the concentrations of reactants, intermediates, and products. The modeling results are then compared against a set of experimental data for validation, and appropriate changes are made, if necessary. Reaction rates are expressed by the modied Arrhenius form, as shown in Equation 2.6. The modied Arrhenius form with temperature dependence is used because non-Arrhenius behavior is often encountered in the temperature ranges encountered in combustion processes [36]. The specic reaction rate constants are best obtained from direct experimental evaluations within dened temperature ranges. A list of elementary combustion reactions and their reactions rates are given by Westbrook and Dryer [36]. CH4 + 2O2 CO2 + 2H2 O Ea RT (2.5)

kA = A (T /Tref )n exp where k is the reaction rate A is the Arrhenius constant T is temperature

(2.6)

Chapter 2. Background Research Tref is the reference temperature (usually 298 K or 25 C) n is the pre-exponential factor Ea is the activation energy R is the ideal gas constant

20

The process for developing and validating chemical kinetic models was outlined by Frenklach et al. [35]: 1. Generate a complete list of elementary reactions. 2. Determine reaction rate constants (see Equation 2.6) for each reaction using literature sources or estimation, paying attention to temperature and pressure dependencies. Provide thermodynamic data to calculate equilibrium reverse rate constants. 3. Conduct controlled experiments that can be used to validate the reactions and rate parameters given in the model. 4. Solve the reaction mechanism kinetics and transport equations using a computer simulation of the experimental conguration. Conduct a sensitivity analysis to determine the impact of specied rate constants on the nal result. 5. Compare the experimental data to the model predicted values. Optimize reaction rate parameters that have the greatest impact on tting desired experimental values. In this study, the user-dened chemical kinetic model was validated against an opposed ow diusion ame setup. The model was numerically solved using the computer application CHEMKIN 4.0 [37]. More information on the experimental conguration and modeling procedure is given in Sections 3 and 4.

2.5.2

Oxidation of Hydrocarbons

A clear and simple explanation of the oxidation of hydrocarbons is given by Glassman [38]. Oxidation reactions are driven by the formation of highly reactive O, OH, and H radicals. During combustion, fuels are oxidized by a series of chain reactions which can be categorized as one of following: 1.) chain initiating, 2.) chain propagating and chain branching, or 3.) chain terminating. Chain initiating occurs when radical species

Chapter 2. Background Research

21

are produced by dissociation of the reactants. The chain is propagated and branched as radicals react with stable compounds to form additional radical species. Finally, the chain terminates when two radicals recombine to form stable species. The following subsections describe the major reaction pathways for the oxidation of alkanes and alkenes under low and high temperature conditions. Very complete mechanisms would contain a number of minor reactions; however, for the sake of simplicity, they have not been included here. Intermediate and Low Temperature Oxidation of Hydrocarbons The oxidation of hydrocarbons is dierent at low temperatures than high temperatures. The general mechanism for the low temperature oxidation of hydrocarbons was rst developed by Semenov [39]. Benson [40] introduced the isomerization reaction of large hydrocarbons (Equation 2.10) to the Semenov mechanism. The following is a simplied form of the Semenov mechanism including isomerization of large hydrocarbons: RH + O2 R +HO2 (2.7)

R +O2 alkene + HO2

(2.8)

R +O2 RO2

(2.9)

RO2 ROOH

(2.10)

ROOH RO +OH

(2.11)

HO2 +RH H2 O2 + R

(2.12)

H2 O2 + M OH +OH +M

(2.13)

The chain is initiated by low temperature oxidation of the hydrocarbon to form an alkyl radical and a hydroperoxy radical (Equation 2.7). Next, the chain is propagated by one of two parallel reactions between alkyl radicals and oxygen to form alkenes, HO2 , and RO2 (Equations 2.8 and 2.9). These reactions compete with each other depending on

Chapter 2. Background Research

22

the temperature. At temperatures above 500 K, Equation 2.8 predominates, wherein the oxygen abstracts a hydrogen from the alkyl radical to form an alkene and a hydroperoxy radical. At temperatures below 500 K, Equation 2.9 is favored, wherein the oxygen adds to the alkyl radical to form an alkylperoxy radical. At low temperatures (below 500 K), propagation continues by the isomerization of RO2 to produce hydroperoxyalkyl radicals (ROOH) (Equation 2.10). The radical pool then builds up by degenerate branching of ROOH to form RO and OH radicals 2.11. Further developments on this low temperature mechanism have been published by Zhao et al. [41]. At intermediate temperatures (above 500 K), the HO2 radical is more abundant, so the reaction is propagated by hydrogen abstraction on the hydrocarbon by HO2 to form hydrogen peroxide (H2 O2 ) and an alkyl radical (Equation 2.12). As the temperature increase, hydrogen peroxide to decompose to form two hydroxyl radicals (Equation 2.13). The fuel-air mixture explodes once the radical pool builds up, and then high temperature oxidation predominates.

High Temperature Oxidation of Alkanes The high temperature oxidation of alkanes larger than methane is initiated by the breaking of CC bonds to form hydrocarbon radicals, as shown in Equation 2.14.

RH + (M ) R +R +(M ) where RH is an alkane molecule R and R are alkyl radicals such as CH3 , C2 H5 , etc. (M) is a non-reacting collision partner

(2.14)

This step dominates because CC bonds are weaker than CH bonds in the molecule. However, CH bonds can also be broken at higher temperatures, as shown in Equation 2.15. In addition, the low temperature abstraction initiation step can occur, as shown in Equation 2.16.

RH + (M ) R +H +(M )

(2.15)

Chapter 2. Background Research

23

RH + O2 R +HO2

(2.16)

The production of H radicals in the chain initiating steps leads to chain propagation, wherein H radicals react with O2 to form OH and O radicals, as shown in Equation 2.17. Chain propagation can also occur by Equations 2.18 and 2.19. The disappearance of fuel occurs after the reservoir of H, OH, and O radicals has been built up, as shown in Equation 2.20. H +O2 O +OH (2.17)

O +H2 OH +H

(2.18)

O +H2 O OH +OH

(2.19)

RH + X R +XH where X is any radical specie, usually O, OH, H, and CH3

(2.20)

Relative rate coecients for H abstraction by radicals from tertiary, secondary, and primary CH bonds are given in Table 2.4 [38]. Tertiary CH bonds are those on a carbon atom connected to three other carbon atoms. Secondary CH bonds are on a carbon atom connected to two other carbons. A primary CH bond is one on a carbon connected only to one other carbon, such as the carbon at the end of a hydrocarbon chain. The table indicates that tertiary CH bonds are the weakest and primary CH bonds are the strongest. If the fuel is an alkene then radicals will abstract H from carbon atoms that are singly bonded because CH bonds on doubly bonded carbon atoms are very strong. The alkane radical then decays to form an alkene and a radical specie, as shown in Equation 2.21. For example, the isopropyl radical, obtained from H abstraction of the secondary CH bond in propane, will decay to propene and a H atom, as shown in Figure 2.4. The process by which the alkyl radical decomposes is called -scission. In -scission, the bond once removed from the radical site will break to form an alkene without a hydrogen shift. Furthermore, CC bonds are more likely break than CH bonds since CC bonds are weaker.

Chapter 2. Background Research

24

Table 2.4: Relative magnitudes of rate constants for H abstraction from dierent CH bonds [38]

R + (X) alkene + R +(M ) where R is a hydrocarbon radical or H atom

(2.21)

Figure 2.4: Decay of isopropyl radical

Oxidation of Alkenes Alkane oxidation ends with the formation of alkenes and a pool of radical species, so the oxidation of these alkene compounds will now be discussed, taking ethene as an example. First, the C=C double bond is attacked primarily by the biradical O , which forms an intermediate species that subsequently decays, as shown in Figure 2.5. Thus, the two primary addition reactions are shown in Equations 2.22 and 2.23. Some minor reactions involving H abstraction by OH and H radicals also play a role, as shown in Equations 2.24 and 2.25, respectively.

Chapter 2. Background Research

25

C2 H4 + O CH3 +HCO

(2.22)

C2 H4 + O CH2 +CH2 O

(2.23)

Figure 2.5: Addition of O radical to ethene from [38]

C2 H4 + OH C2 H3 +H2 O

(2.24)

C2 H4 + H C2 H3 +H2

(2.25)

Then the vinyl radical (C2 H3 ) decays to acetylene, as shown in Equation 2.26. The acetylene is consumed by a reaction with the biradical O to form a methylene radical and carbon monoxide, as shown in Equation 2.27. The fate of CH3 , CH2 O (formaldehyde), CH2 , and CO are described in the mechanism for methane [38]. C2 H3 +M C2 H2 + H +M (2.26)

C2 H2 + O CH2 +CO

(2.27)

2.5.3

Oxidation of Methyl Esters

Oxidation of Methyl Acetate Dagaut et. al [42] developed a chemical for the oxidation of methyl acetate and validated it against experiments conducted in a jet-stirred reactor at 800-1230 K and 10 atm.

Chapter 2. Background Research

26

Methyl acetate is a methyl ester with two carbon atoms in the hydrocarbon chain, as shown in Figure 2.6. The study indicates that methyl acetate is primarily consumed by H abstraction by H, O, and OH radicals, as shown in Equations 2.28 and 2.29.

Figure 2.6: Structure of Methyl Acetate

CH3 C(= O)OCH3 + X CH3 C(= O)OCH2 + XH

(2.28)

CH3 C(= O)OCH3 + X CH2 C(= O)OCH3 + XH where X is an H, OH, or O radical

(2.29)

The resulting methyl acetate radicals are then decomposed by -scission, as shown in Equations 2.30 and 2.31. CH3 C(= O)OCH2 CH3 C(= O) + H2 C(= O) (2.30)

CH2 C(= O)OCH3 CH2 C(= O) + CH3 O

(2.31)

H abstraction on CH2 C(=O) produces HCCO, which further reacts with O2 to form CO2 . Thermal decomposition reduces CH3 C=O to CH3 and CO, and CH3 O to CH2 O (formaldehyde) and H. Further decomposition of formaldehyde was found to lead to the formation of the main active species in the mechanism: HCO, H, OH, HO2 , and O. Oxidation of Dimethyl Carbonate A chemical kinetic model for dimethyl carbonate in an opposed ow diusion ame was studied by Glaude and coworkers [43]. Dimethyl carbonate ((CH3 O)2 C=O) has a chemical structure similar to that of the FAMEs found in biodiesel. In addition, the

Chapter 2. Background Research

27

oxidation of DMC leads to the formation of the methoxy formal radical (CH3 OC =O), which is structurally similar to the key moiety (ROC =O) found in all oxygenated fuels, including FAMEs. Figure 2.7 shows the major reaction pathways for the consumption of DMC in their model. Initially, DMC is consumed by reaction with H and OH to produce the DMC radical. This intermediate quickly decomposes to CH2 O (formaldehyde) and CH3 OC =O (methoxycarbonyl radical). Finally, 78% of the CH3 OC =O radical leads to CO2 , while the remaining 22% forms CO.

Figure 2.7: Main reaction pathways for DMC in the opposed ow diusion ame (percentages refer to a reaction pathways share of a species consumption) from [43]

A key contribution of this paper is new reaction rate constants for the decomposition of CH3 OC =O. The authors state that these reaction rate rules can be applied for any ROC =O radical, including the ones found in FAMEs. The study also concludes that the formation of the (ROC =O) radical leads to more CO2 than CO by decomposition. From a soot formation standpoint, the formation of CO2 is not desirable because two oxygen atoms are bonded to one carbon atom. Ideally, each fuel-bound oxygen atom

Chapter 2. Background Research

28

should bond with one carbon atom to suppress the formation of carbon-carbon bonds, which lead to soot. Other studies on the oxidation of esters have provided additional evidence supporting this theory [44].

Oxidation of Methyl Butanoate Methyl butanoate (nC3 H7 C(=O)OCH3 ) is a short chain FAME with a structure very similar to the long chain FAMEs found in biodiesel, which have the general formula RC(=O)CH3 (where R is an alkyl or alkenyl radical). Fisher et al. [8] have studied the oxidation of methyl butanoate in an eort to better understand the oxidation of the long chain FAMEs. The authors state, methyl butanoate was chosen as a surrogate molecule for the larger methyl esters, in order to obtain a reaction mechanism of manageable size. Yet methyl butanoate is large enough to allow fast RO2 isomerization reactions important in the low-temperature chemistry that controls fuel autoignition under conditions found in diesel engines. Figure 2.8 displays the key low temperature and high temperature pathways for the oxidation of methyl butanoate. Initially, the fuel molecule undergoes H abstraction to form one of four possible radicals. In this case, H is abstracted from the carbon alpha to the carbonyl group (reaction 1). The resulting radical species can then follow one of two pathways depending on the temperature. At high temperatures the radical decomposes by -scission to produce one stable molecule and a CH3 radical (reaction 2). On the other hand, at low temperatures the radical reacts with O2 to produce a complex radical (reaction 3). An isomerization reaction then occurs to shift a hydrogen to the O2 addition products (reaction 4). This species can then follow one of two reaction pathways. In this rst instance, the molecule decomposes unimolecularly to form an OH radical, ethylene, and an epoxide-like species (reaction 5). In the second instance, O2 is added to the molecule (reaction 6), which leads to a series of chain branching reactions (reactions 7 and 8). An attempt to validate the mechanism was made by testing it against existing combustion data, consisting of pressure measurements. However, the experimental data was not well characterized for model validation. Therefore, the authors urged researchers to obtain more complete experimental data so that the mechanism could be rigorously tested.

Chapter 2. Background Research

29

Figure 2.8: Main reaction pathways for the oxidation of methyl butanoate from Fisher et al. [8]

2.6

Mechanism of Soot Formation in Combustion Processes

The term soot refers to tiny amorphous carbon particles produced from the combustion of hydrocarbon fuels. Soot emissions have become an environmental concern due to the negative impacts of particulate matter on the human respiratory system. In addition, soot particles in the atmosphere can contribute to global warming by altering the radiative balance of the atmosphere [45, 46]. The presence of soot particles in hydrocarbon ames is identiable by their characteristic yellow-orange appearance. This color is generated by photons emitted from the solid carbon particulates. On the other hand, ames that do not contain soot are

Chapter 2. Background Research

30

blue in color. The formation of soot particles in a combustion environment is a highly complex mechanism and depends on a number of factors, e.g. fuel composition, temperature, fuel-oxygen ratio, ame conguration, etc.. However, the formation of soot has been shown to be highly controlled by chemistry related phenomenon [47]. Glassman [38] has summarized the work of other researchers to describe the dominant route of soot formation. Initially, the fuel breaks down to acetylene, as mentioned in the previous discussion on hydrocarbon oxidation. In the high-temperature post-ame regime, soot formation is initiated by the growth of small straight-chain alkenes (acetylene) to small aromatic compounds (e.g. benzene). The aromatic hydrocarbons then react sequentially with smaller hydrocarbons (acetylene, in particular) to form larger polyaromatic hydrocarbon (PAH) species. Gaseous PAH molecules continue to nucleate until the smallest identiable soot particles appear, with diameters of a few nanometers. Figure 2.9 [38] shows the mechanism for soot formation in more detail. Initially, the acetylene (C2 H2 ) undergoes H abstraction to form the vinyl radical. The vinyl radical then reacts with another acetylene molecule to form the 1,3-butadienyl radical. The 1,3-butadienyl radical can also be readily formed from C4 hydrocarbons via hydrogen abstraction then -scission. In diusion ames, the alternate route A is then followed, wherein the 1,3-butadienyl radical reacts again with acetylene to form the cyclic phenyl (C6 H5 ) radical, following ring closure. The phenyl radical is essentially a benzene molecule missing one hydrogen atom. The phenyl radical is also produced by alternate route C, wherein methyl acetylene (propyne C3 H4 ) pyrolyzes rapidly to form the aromatic. The phenyl radical can proceed to grow into a larger aromatic via the two-step hydrogen abstraction-carbon addition (HACA) mechanism. In the HACA mechanism, the aromatic molecule is converted to a radical by hydrogen abstraction, and then grows in size by the (carbon) addition of an acetylene molecule. The HACA mechanism continues until larger PAH molecules appear. As the concentration of gaseous PAH species increases, nucleation occurs and soot particles begin to appear. This thesis study does not directly measure the soot levels generated in biodiesel diusion ames. However, the concentrations of many soot precursors are measured, specically, acetylene, C4 hydrocarbons, 1,3-butadiene, and benzene. Analyzing the concentrations of these species in the diusion ame can oer an insight into the sooting potential of various biodiesel fuels.

Chapter 2. Background Research 31

Figure 2.9: General mechanism for soot formation from Glassman [38]

Chapter 3 Experimental Apparatus and Analytical Methodology


The details of the experimental setup and the analytical techniques employed are discussed in this chapter. Figure 3.1 is a schematic of the setup. The purpose of the setup was to generate an opposed ow diusion ame from the liquid fuel feed stock (i.e. methyl butyrate or methyl crotonate). The concentrations of stable species and the temperature prole in the ame were then obtained. The setup consists of the fuel delivery system, the opposed ow diusion ame burner, the sample collection apparatus, and a number of analytical instruments. The fuel delivery system pumped the liquid fuel into a vaporizing unit that produced a gaseous fuel mixture. The gaseous fuel stream and the oxidizer stream then owed into an opposed ow burner to produce a planar ame. Samples were extracted from the ame region using a quartz micro-probe connected to a vacuum pump. The hydrocarbon species and several oxygenated compounds were analyzed by a gas chromatograph (GC) coupled with a ame ionization detector (FID). Carbon dioxide and carbon monoxide were quantied by a non-dispersive infra-red (NDIR) analyzer. Carbonyl compounds were measured by a high pressure liquid chromatograph (HPLC) coupled with a ultraviolet-visible (UV-Vis) spectrophotometer. The ame temperature was measured by an R-type thermocouple.

32

Chapter 3. Experimental Apparatus and Analytical Methodology

33

Figure 3.1: Schematic of the experimental setup

Chapter 3. Experimental Apparatus and Analytical Methodology

34

3.1

Opposed ow diusion burner setup

Opposed ow diusion ames oer greater experimental and modeling advantages than co-ow ames, although their stability is more sensitive to ow conditions [38]. The opposed ow conguration results in a at ame which simplies ame analysis to a onedimensional system. Therefore, the species concentration and temperature are a function of axial distance only. Furthermore, the modeling of uid mixing is simplied in laminar ames because fuel and oxidizer mixing is limited to diusional processes (i.e. turbulent mixing patterns are not considered). The experiments utilized two identical at ame burners1 with circular burner ports. The two burner ports are placed opposite each other in the same vertical plane. Fuel was fed through the bottom port while oxygen was fed through the top. The two opposing streams ow into each other to create a stagnation plane between the two ports. The vertical location of the stagnation plane depends on the momentum of the two streams. The stagnation plane prevents any non-diusional mixing between the fuel and oxidizer. The mixture was ignited to create at stoichiometric ame lying slightly above the stagnation plane. The exact location of the ame front depends on the mass fraction of fuel and oxidizer in the fuel stream and oxidizer stream. A diagram of the burner port is shown in Figure 3.2. Each burner port consists of a stainless steel housing enclosing a porous sintered bronze matrix. The porous material is divided into inner and outer coaxial cylinders of diameter 25.4 mm and 38.1 mm, respectively. The inner cylinder directs the fuel or oxidizer streams towards each other, and the outer annulus can be used to create a nitrogen shroud around the ame for minimizing external ow disturbances. The annulus shrouding feature was not exploited in this study. The temperature of the gases owing through the ports is controlled by circulating water through porous plugs packed between two co-axial cylinders. This ensures that the gases have a uniform laminar ow and at velocity prole at the ports surface. The water is maintained at 70 C and circulated by a heating recirculator2 . Condensation of fuel vapor in the bottom port is prevented by maintaining the port temperature near 80 C with heating tapes3 .

1 2

Purchased from Holthuis & Associates McKenna Flat Flame Burners PolyScience Heating Recirculator Model 210 3 Omega R FGS Standard Insulated High Temperature Heating Tapes

Chapter 3. Experimental Apparatus and Analytical Methodology

35

Figure 3.2: Diagram of burner port

A photograph of the actual burner setup is shown Figure 3.3. The two burners are coaxially mounted 20 mm apart facing each other. A custom-built aluminum holder provides support to the burners, and a clear quartz shroud 22 cm in diameter protects the ame from airow disturbances from the surrounding environment. The entire assembly is mounted atop a Newport translation stage which moves along the vertical axis with the rotation of a micrometer knob. This made it possible to obtain a vertical prole of the ame temperature and emissions between the two burner ports.

Chapter 3. Experimental Apparatus and Analytical Methodology

36

Figure 3.3: Photograph of burner setup

3.2

Fuel Preparation and Vaporization

The fuel stream fed to the burner was a mixture of the biodiesel surrogate fuel (i.e. methyl butanoate or methyl crotonate) and nitrogen gas. Properties of each surrogate fuel are shown in the following Table 3.1, which indicates that both fuels are liquids at room temperature. Excess dissolved gases were eliminated by degassing the fuel in an ultrasonic bath4 . The fuel was heated in the water bath to 70 C, and then ultrasonic sound waves were applied until tiny bubbles no longer rose to the surface of the liquid5 . The fuel was then pumped from a bottle to an ultrasonic atomizer. The atomizer probe sprays the fuel into a stainless steel mixing chamber where it mixes with preheated nitrogen gas. The mixture of fuel and nitrogen was then fed to the bottom burner port. The fuel was pumped from its stock bottle by a peristaltic pump head driven by a motor6 . The pumping system was calibrated for each fuel at the beginning of each
4 5

Cole-Parmer 8890 Ultrasonic Cleaner Approximately 60 minutes 6 Cole-Parmer Masterex R L/S Microprocessor Pump Drive

Chapter 3. Experimental Apparatus and Analytical Methodology

37

Fuel Compound

Boiling point range (degrees C)

Purity (%) 99 98

Specic gravity

Methyl Butanoate Methyl Crotonate

100-103 116-120

0.898 0.945

Table 3.1: Physical properties of the fuels used

experiment. This involved using a graduated cylinder to measure the volume of liquid pumped, and then setting the pump to correspond with the measured volume. Initially, a polycarbonate pump head housing and Viton R tubing were used to dispense the fuel; however, it was found that methyl butanoate was reacting with the materials and causing component failures. This problem was overcome by replacing the damaged components with parts made of PTFE7 . PTFE is inert to methyl butanoate and methyl crotonate. The pumped fuel was delivered to a ultrasonic atomizer8 unit that breaks the fuel into micro-droplets. The units ultrasonic power supply converts 60 Hz energy to a high frequency electrical energy at 40 kHz. Then, a piezoelectric transducer converts the electrical energy to mechanical vibrations. The vibrations are intensied in a probe and focused at its tip. The liquid fuel is dispensed in the probe, where it spreads out as a thin lm on the tip. The oscillations at the tip atomizes the liquid into micro-droplets to form a gentle, low viscosity mist. The median droplet size is 45 microns [48]. A picture of the fuel vaporization unit is shown in Figure 3.4. The atomized fuel is sprayed into a mixing chamber, where it mixes with nitrogen gas9 . The nitrogen was preheated by an inline electric process heater10 , and then delivered to the mixing chamber via 1/4 inch copper tubing. The temperature in the chamber and all copper tubing was maintained at 85 C using heating tapes11 . The high temperature in the mixing chamber served to vaporize the fuel droplets. The temperature inside the chamber is measured using a stainless steel sheathed K-type thermocouple inserted at the top of the chamber.
Polytetrauoroethylene or Teon R Sonics VibracellTM VC 134 AT 9 Grade 4.8 10 Omega R process heater 11 Omega R FGS Standard Insulated High Temperature Heating Tapes
8 7

Chapter 3. Experimental Apparatus and Analytical Methodology

38

The mixing chamber is a custom-made stainless steel column12 30 cm in length and 6 cm in outer diameter. The lower half of the column is packed with a 15 cm long bed of glass marbles. The bed increases the path length traversed by gaseous fuel and nitrogen molecules; thereby, increasing mixing of the two streams.

Figure 3.4: Photograph of fuel vaporization column

Table 3.1 indicates that neither fuel used in these experiments is a pure component. In fact, each fuel contains 1-2% of viscous substances which would not vaporize in the mixing chamber. These greasy viscous substances were found to clog the burners sintered porous matrix. Therefore, an inline coalescing lter13 was installed at the exit of the mixing chamber to remove these impurities. The lter assembly consists of a lter element, an aluminum housing, and an EPDM14 o-ring. The lter element is made of borosilicate glass micro-bers bound by a silica inorganic resin. The lter coalesces and removes 99.5% of 0.1 micron sized particles. The puried gaseous mixture of fuel and nitrogen ows to the bottom port of the burner apparatus via a heated15 1/4 copper tube.
12 13

Designed and constructed at Dept. of Mechanical and Industrial Engineering Machine Tool Lab Manufactured by United Filtration Systems 14 Ethylene Propylene Diene Monomor 15 Omega R FGS Standard Insulated High Temperature Heating Tapes

Chapter 3. Experimental Apparatus and Analytical Methodology

39

3.3

Supply of Fuel and Oxidizer Streams

The ow rates of fuel and oxidizer streams through the burner ports were key parameters in these experiments. It is important that the momentums of the two streams be nearly equal, so that a stagnation plane is created where the streams meet. The molar concentrations of fuel and oxidizer in each stream needs to be sucient enough to light a ame; however, high sooting ames are not desired. The ow rates of nitrogen, oxygen and air were controlled using rotameters, while the ow rates of liquid fuel were controlled by a peristaltic pump. The inlet oxidizer and fuel stream concentrations were selected based on the following criteria: i.) a low sooting ame is preferred to prevent clogging of sampling probe; ii.) a very hot ame that will damage the probe is unwanted; iii.) a balanced momentum of the two streams is required to form a stagnation plane at their intersection; iv.) excessive unburned fuel is not desired; v.) at the ame plane, an N2 /O2 ratio near that of air is desired to make the study relevant to actual ames. The molar composition of the fuel stream was 4.72% fuel (methyl butanoate or methyl crotonate) and 95.28% nitrogen. Additional ow parameters are given in Appendix A. The volumetric ow of fuel to the vaporizing column was chosen to obtain an identical molar ow rate for both fuels. Since the two fuels have dierent liquid densities, the volumetric ow rates were dierent. Methyl crotonate was dispensed at 0.67 ml/min, while methyl butanoate was dispensed at 0.72 ml/min. In our experiments, the temperature of the fuel stream exiting the bottom burner port was 80 C. The molar composition of the oxidizer stream was 57.75% nitrogen and 42.45% oxygen16 . Unexpectedly, the temperature of the unheated oxidizer stream exiting the top burner was near 141 C. The cause of this increased temperature was heat convected towards the top burner port from rising combustion product gases.

3.4

Gas Sampling System

The previous sections discussed the procedure for producing an opposed ow diusion ame. Once the ame was generated, a sampling system was used to obtain qualitative and quantitative information about the ames characteristics. Specically, the species concentrations at various points between the two burner ports was measured to obtain characteristic proles. The following sections discuss the gas sampling system, as well as
16

Grade 4.3

Chapter 3. Experimental Apparatus and Analytical Methodology the sampling procedure.

40

3.4.1

Sampling Apparatus

Microprobes, due to their small perturbation of ow elds, are commonly used in ame studies to acquire the concentration of stable species [49, 50, 51, 52]. The gas sampling system in these experiments consists of a quartz microprobe connected to a dual-stage pump17 via 1/4 inch PTFE tubing, a vacuum pressure gauge, and an inline lter. The microprobe is mounted on a sliding stage, allowing it to move into and out of the ame region easily. A schematic of the microprobe is shown in Figure 3.5. The rst stage (vacuum) of the pump creates a suction in the sampling line to withdraw gas samples from the ame. An analytical instrument was connected downstream of the second stage (compressor) of the pump to study the gases owing through the line. The compressor head on the pump pushes samples into the analytical instrument. A two-way valve, installed downstream of the pump, controls gas ow to the instrument.

OD: 585 m ID : 202

3.75

6.35

300 mm (approx.)

Figure 3.5: Schematic of microprobe (not to scale)

Previous studies have identied precautions to take when using the microprobe sampling technique. The primary objective is to eliminate chemical reactions within the probe and sampling lines. A combination of rapidly reducing temperature and pressure in the probe helps meet this objective. Kassem and coworkers [53] studied the eect of microprobe cooling on fuel-rich, laminar at ames of chlorinated hydrocarbons. Their results indicate that cooled probes,
17

KNF oil-free dual-staged pump Model UN035.3 ST11 with heated heads

Chapter 3. Experimental Apparatus and Analytical Methodology

41

as opposed to uncooled probes, provided more accurate proles of species concentration. Schoenung and Hanson [49] showed that carbon monoxide (CO) measurements in the post-ame region of a premixed methane/air ame were aected by the pressure within the probe and sampling lines. Their results indicate that the CO concentration increases as the pressure in the probe decreases, with the concentration reaching equilibrium around 50 mm Hg. This nding suggests that CO is converted to CO2 in the probe region unless the pressure is 50 mm Hg or below. Therefore, low temperatures and pressures in the probe are required to quench the reactions. Fristrom and coworkers [54] argue that it is not rapid temperature drop that is required for successful ame sampling. Instead, a combination of rapid pressure drop and the destruction of radicals on the probe walls is responsible for quenching reactions at the probe tip. They explain that 2nd order molecule-radical reaction rates vary with the square of the gas density, which varies linearly with pressure; therefore, the reaction rate decreases with a decrease pressure. Furthermore, they argue that if a rapid temperature drop is induced while keeping pressure constant, then reactions with activation energies of 20 kJ/mol or lower will increase in rate. These reaction rates vary quadratically with density, which varies inversely with temperature at constant pressure; therefore, decreasing the temperature alone would not quench all ame reactions. The experiments in this study used a quartz microprobe fabricated in the Department of Chemistry Glass Blowing Shop. The inner diameters (ID) of the probes tip could not be precisely controlled; therefore, a number of probes were made and measurements were taken to select the best one. The probe tip ID was measured using a travelling microscope mounted on a metric scale. A previous study by Syed [9] determined the appropriate probe tip size. The study measured CO2 concentrations in a propane-air ame using probe tips with various IDs. Syed suggests that a probe tip with an ID of approximately 180-220m is ideal. Probes with larger IDs do not successfully quench reactions at the probe tip. In addition, probes with smaller IDs restrict gas ow through the sampling line, such that it is not possible to take measurements. As mentioned previously, the microprobe was connected to a doubled-headed pump via 1/4 inch PTFE tubing. The vacuum pump head creates a vacuum pressure of 710-730 mm Hg18 , which is sucient to quench most reactions. The probe was not cooled since previous studies have shown that a rapid pressure drop alone provides accurate sampling.
18

Gauge vacuum pressure measured by a vacuum pressure gauge

Chapter 3. Experimental Apparatus and Analytical Methodology

42

Extra precautions were taken to eliminate leakage of gases from the surrounding environment into the sampling line. Leaks into the sampling system dilute the sample gas and lead to incorrect measurements. For this reason, Swagelok R ttings are used for all connections. Leaks into the sampling line were detected by using a container lled with dry ice (i.e solidied carbon dioxide). The container was placed near a suspected point of leakage, and carbon dioxide gas lls the surrounding area. The NDIR analyzer (see section 3.5.1) was connected to the sampling line and the dual-stage pump was turned on. If a spike in CO2 concentration was observed on the analyzer, then there was a leak present in the sampling line. The necessary steps were then carried out to eliminate the leak (e.g. the ttings were changed, the tube was replaced, etc.).

3.4.2

Sampling Procedure

The sampling objective was to obtain ame measurements at various points along the vertical axis separating the two burner ports. Thus, the sliding stage, on which the microprobe is mounted, was inserted between the two burner ports. A window was cut into the quartz shroud enclosing the burner setup to permit insertion of the sampling probe. Figure 3.6 is a schematic of the probe and burner setup. The tip of the probe was placed approximately 1.5 mm behind the central vertical axis separating the burner ports. This allows samples to be withdrawn from the middle points of the ame region. After insertion, the probe was held stationary, while the burner assembly was moved along the vertical axis with the turn of a micrometer knob on the translation stage. One complete turn of the micrometer knob corresponds with a 0.5 mm change in distance. Clockwise rotation moves the burner assembly up. A measuring system was used to dene the exact position of the microprobe between the two burner ports. For this purpose, the bottom (fuel) port is taken as the zero height, while the top (oxidizer) port is taken as the maximum height. The position of the probe was zeroed by touching its tip to the bottom port. The reading on the micrometer knob was thus noted as the zero distance. Each counterclockwise turn of the micrometer knob moved the burner assembly down; thus, increasing the distance between the probe and the fuel port, i.e., moving the probe upwards. The exact height at which the probe withdraws samples is equal to the total distance travelled plus the outer radius of the probe. Once the probe was positioned at the desired height, the dual-stage pump was turned

Chapter 3. Experimental Apparatus and Analytical Methodology

43

Oxidizer Port Probe tip behind central vertical axis Microprobe Burner moves Vertically

Sliding Stage Fuel Port Microprobe moves Horizontally Burner Platform with rotating micrometer

Figure 3.6: Schematic of microprobe and burner setup (not to scale)

on and gases withdrawn from the ame lled the sampling line. The sampling lines were purged before any analytical measurements are taken. The purge time varies depending on the location of the probe in the ame. Sampling points away from the center of the ame require short purge times (e.g. 2-3 minutes), since the low temperature, high density gases permit high owrates. In contrast, sampling points near the center of the ame require longer purge times (e.g. 15 minutes), since gases in this high temperature region have an extremely low density and permit low owrates. The sampling line lled with gases from the specied ame region were then ready for analysis.

3.5

Analytical Tehniques

The hydrocarbon species and several oxygenated compounds (e.g. methanol, methyl butanoate, acetone, etc.) were analyzed by gas chromatography coupled with a ame ionization detector (GC/FID). Carbon dioxide and carbon monoxide were quantied by non-dispersive infra-red (NDIR) analysis. Carbonyl compounds (e.g. formaldehyde, acetaldehyde, and acrolein) were measured by high pressure liquid chromatography coupled with an ultraviolet-visible spectrophotometer (HPLC/UV-Vis). The ame temperature was measured by an R-type thermocouple.

Chapter 3. Experimental Apparatus and Analytical Methodology

44

3.5.1

Non-Dispersive Infrared Analysis

Non-dispersive infrared (NDIR) analysis is a technique used to measure gas concentrations based on the energy absorption characteristics of a gas in the infrared red region. Figure 3.7 is a schematic of a typical NDIR instrument [55]. The NDIR instrument passes infrared (IR) light through two identical tubes, in parallel, and then onto a detector. The rst tube is lled with nitrogen, which does not absorb IR light and serves as a reference cell. The second tube contains the sample gas, which absorbs IR energy. The detector measures the dierence in energy between the two streams of IR light [55]. This dierence is the absorption, which is proportional to the concentration of sample gas by the Beer-Lambert Law in Equation 3.1:

A= bc

(3.1)

where A is the gas absorbance. is the molar extinction coecient (concentration1 length1 ). b is the path length that the beam travels in the sampling tube. c is the gas concentration.

Figure 3.7: Schematic of NDIR instrument from [55]

Chapter 3. Experimental Apparatus and Analytical Methodology CO and CO2 Measurements

45

The NDIR instrument19 was used to quantify levels of CO and CO2 in the ame samples. The instrument is capable of measuring concentrations from 0% to 10.5%. It contains an upper and lower screen displaying the concentration of CO and CO2 , respectively. Initially, the instrument was zeroed and calibrated. To zero the instrument, the built-in pump was started to allow air ow through the unit; then the zero adjustment knob was turned to set the display to zero. Next, the unit was calibrated with a gas mixture containing 8.9% CO and 9.1% CO2 . The gas mixture was passed through the detector, and the span adjustment knob was turned to match the calibration gas concentration. Flame measurements were taken by using the dual-stage pump to withdraw samples from the ame and then pushing them into the NDIR analyzer. The analyzers builtin pump was not used during sampling. The concentrations were recorded once the displayed reading remained constant for 5 minutes. As mentioned previously, samples withdrawn from points away from the ame center have higher owrates than samples from within the middle of the ame. Therefore, the analyzers display reading stabilizes much quicker for the former (e.g. 1-3 minutes) than the latter (e.g. 15-20 minutes).

3.5.2

Gas Chromatography

Gas chromatography (GC) is a technique used for separating volatile organic compounds based on their dierences in partitioning between a owing mobile phase and a stationary phase. The GC method was used to measure C1 - C5 alkanes, alkenes, and alkynes, benzene, methyl butanoate, methyl crotonate, methanol, and acetone. A GC instrument consists of a owing mobile phase (carrier gas), a stationary phase (separation column), an injection port, an oven, and a detector [56]. A schematic of the GC instrument is shown in Figure 3.8. The carrier gas carries the sample gas through the separation column, and compounds are separated due to partitioning between the two phases. Since partitioning behavior is a strong function of temperature, the separation column is placed inside a temperaturecontrolled oven. Separation of compounds with a range of boiling points is achieved by starting at low temperatures and then increasing the temperature until high boiling point compounds are eluted. The injection port is always maintained at a temperature higher
19

NOVA NDIR Analyzer Model 7800P2A

Chapter 3. Experimental Apparatus and Analytical Methodology


Injector Detector

46

Data Acquisition System

Carrier Gas (mobile phase)

Column Oven

Separation Column (stationery phase)

Figure 3.8: Schematic of GC instrument

than the boiling point of the least volatile compound in the mixture [56]. A 10% sorbitol pre-column is installed prior to the separation column to remove moisture and oxygen contaminants from the sample gas. The amount of time a given component spends in the separation column is called the retention time. The retention time of a given component remains the same provided the mobile phase, stationary phase, temperature control, and gas owrates remain constant. As each component of the separated sample falls onto the detector a quantitative response in the form of a peak is generated. A series of peaks with the retention time on the x-axis and the detector (e.g. voltage) on the y-axis is called a chromatogram (see Appendix C for an example). The peak retention time is used to identify each compound, and the peak area is used to determine the quantity of the compound. The instrument used in these experiments was a Varian 3800 GC20 with a 1079 injector and a ame ionization detector (FID). Figure 3.9 shows a schematic of an FID [57]. The FID consists of a hydrogen/air ame and a collector plate. As gases ow from the separation column, the ame burns organic molecules to produce ions. The ions are attracted to the collector plate, which generates a voltage depending on the quantity of ions collected [57]. Additional details of the GC setup are discussed in the follow subsections.
20

Remotely controlled by a PC using STAR Chromatography Worstation 4.5

Chapter 3. Experimental Apparatus and Analytical Methodology

47

Figure 3.9: Schematic of FID from [57]

GC Carrier Gas The carrier gas (mobile phase) used for the GC was 99.997 % helium. Hydrocarbon and oxygen traps were placed before the column to lter out any contaminants. The owrate of helium through the separation column was varied depending on the compounds being studied. Details are available in subsection GC Measurement Procedures below.

Injection System The purpose of the injection system is to load the sample gas onto the separation column. The GC has a 1079 universal capillary injector, which can be run in several modes based on the type of injector insert used. The injector temperature was set at 200 C to prevent condensation of sample components. An unpacked 3.4 mm ID insert for split mode operation was used. Split mode operation is discussed below. Flame samples are pushed into the GC sampling line by the dual-stage pump. A valve downstream of the GC sampling line was initially closed to increase the sample pressure in the line. The sample pressure, measured by an in-line pressure gauge, was allowed to reach approximately 8.7 psia. At this point the dual-stage pump was shut o and the GC sampling line was closed to isolate the sample for injection to the GC column. The ow of sample gas and carrier gas into the column is then controlled by the gas sampling valve (GSV) and the split

Chapter 3. Experimental Apparatus and Analytical Methodology valve.

48

The GSV is a 10 port rotary valve which directs the sample and carrier gases into the injector. Schematics of the GSV are shown in Figures 3.10 and 3.11. The GSV has two positions; the ll position and the load position. In the ll position, the sample gas ows through the 0.25 ml sample loop, while carrier gas ows through the sorbitol pre-column, injector, and separation column. As the GSV turns to the load position, the sample trapped in the loop comes in-line with the carrier gas ow. The sample is carried through the sorbitol pre-column and into the injector. From the injector, it is directed into the column. The GSVs duration in the load position is a user controlled parameter, usually less than a minute. As the GSV returns to the ll position, the normal ow pattern resumes.
Sample in Sample out Sorbitol precolumn 6 7 8 GSV vent Helium To injector 5 4 3 2 9 10 1 Helium Sample loop

Figure 3.10: Schematic of injector in the ll position

Sample in Sample out Sorbitol precolumn 6 7 8 GSV vent Helium To injector 5 4 3 2 1 Helium Sample loop

9 10

Figure 3.11: Schematic of injector in the load position

Chapter 3. Experimental Apparatus and Analytical Methodology

49

The sample gas in the injector can be introduced to the column via split mode or splitless mode. In splitless mode, the entire sample is loaded onto the column. This mode is advantageous for detecting trace compounds in the sample gas. However, highly concentrated samples can damage the separation column. Therefore, split mode operation was used here, wherein only a portion of the sample entered the column. A special 3.4 mm ID glass insert for split mode operation was used in these experiments. In split injection, a solenoid is activated such that only a portion of the sample gas ows into the column while the remainder is vented by the carrier gas owing through the split vent. The split ratio is calculated as the ratio of the carrier gas ow through the column to the carrier gas ow through the split vent (see Tables 3.3 and 3.4 for split ratio values). The ow of carrier gas through the split vent and separation column is manually adjusted by the split ratio ow controller and the back pressure regulator, respectively. Table 3.2 indicates the settings used for the split valve and GSV.
Time Interval (min) 0 0.1 1 Gas Sampling Valve Setting Fill Inject Fill Split Valve Setting Split Split Split

Table 3.2: Settings for the gas sampling valve and split valve

GC Measurement Procedures Two dierent GC measurement procedures were developed to analyze the ame samples. The rst method was designed to measure C1 -C5 hydrocarbons using a HP-Al/S PLOT capillary column21 . The second method measured methyl butanoate, methyl crotonate, methanol, and acetone using a DB-624 column22 . Two dierent columns were used because neither is capable of separating both hydrocarbon and oxygenate species. The PLOT column is ideal for separating hydrocarbon compounds that are gases are room temperature. The DB-624 column is suitable for the separation of oxygenated compounds. The two columns were not mounted in parallel. Instead, rst hydrocarbons
21 22

Agilent Technologies HP-Al/S 50 m x 0.53 mm (L x ID) Agilent Technologies DB-624 30 m x 0.32 mm (L x ID)

Chapter 3. Experimental Apparatus and Analytical Methodology were measured, and then the columns were switched for sampling of oxygenates. Methodology for Analysis of Hydrocarbons

50

The PLOT column has a stationary phase comprised of alumina deactivated with sodium sulfate. The columns length, inner diameter, and lm thickness are 50 m, 0.53 mm, and 15 m, respectively. Additional separation characteristics of the column are shown in Appendix C. A column oven temperature program was designed to obtain quick separation of low molecular weight compounds (e.g. methane, ethane, etc.) and high molecular weight compounds (e.g. benzene). The column oven temperature program is shown in Figure 3.12. The initial temperature is held at 50 C to separate the more volatile species. A gradual increase in temperature (15 C/min) then forces less volatile compounds out the column. Finally, the temperature is held at 200 C to elute the remaining components with very high boiling points. A sample chromatogram using this column and oven temperature program is shown in Appendix C.

Figure 3.12: Oven temperature program for measuring hydrocarbons using HP Al/S PLOT column

The inlet pressure of hydrogen, air, and helium were 40, 60, and 80 psig, respectively. The ow of gases in the GC and FID are shown in Table 3.3.

Chapter 3. Experimental Apparatus and Analytical Methodology


Split Vent Flow (He) Flowrate (mL/min) 18.0 Carrier Gas Flow (He) 5.0 Carrier + Make up (He) 30 Hydrogen for FID 30 Air for FID 300

51

Split ratio = 1 : 3.6

Table 3.3: Flow rate of GC gases when measuring hydrocarbons

Methodology for Analysis of Oxygenates The DB-624 column was chosen for measuring oxygenates because the mid-polarity stationary phase provides good separation of alcohols, ketones, and aldehydes. The columns length, inner diameter, and lm thickness are 30 m, 0.53 mm, and 1.8 m, respectively. Additional properties and separation characteristics of the column are shown in Appendix C. The oven temperature program was designed to separate oxygenated compounds. The column oven temperature program is shown in Figure 3.13. The initial temperature is held at 70 C to separate the lighter, less polar species. A gradual increase in temperature (15 C/min) then forces heavier, more polar compounds out the column. Finally, the temperature is held at 200 C to elute the remaining components. A sample chromatogram using this column is shown in Appendix C The inlet pressure of hydrogen, air, and helium were 40, 60, and 80 psig, respectively. The ow of gases in the GC and FID are shown in Table 3.4.
Split Vent Flow (He) Flowrate (mL/min) 100 Carrier Gas Flow (He) 2.0 Carrier + Make up (He) 30 Hydrogen for FID 30 Air for FID 300

Split ratio = 1 : 50

Table 3.4: Flow rate of GC gases when measuring oxygenates

Chapter 3. Experimental Apparatus and Analytical Methodology

52

Figure 3.13: Oven temperature program for DB-624 Column

GC Calibration Procedure

The GC was calibrated for hydrocarbons using four dierent Scotty R calibration gas mixtures: 100 ppm C1 -C6 alkanes in nitrogen; 100 ppm C2 -C6 alkenes in nitrogen; 15 ppm C2 -C4 alkynes in nitrogen; and 100 ppm benzene in air. The calibration was performed by owing each gas mixture directly into the GC sampling line. The operating conditions for the GC were identical to those aforementioned. GC calibration for the oxygenated species involved using the fuel delivery and burner setup because gas cylinders were not available. The liquid oxygenate was pumped into the vaporization and diluted with nitrogen gas. The exact concentration of the nitrogenoxygenate mixture was determined from the nitrogen gas owrate and the liquid ow rate into the mixing column. This calibration gas was owed to the bottom burner port, and samples were obtained using the microprobe sampling technique. The aforementioned GC sampling method was used to calibrate the system.

Chapter 3. Experimental Apparatus and Analytical Methodology

53

3.5.3

High Pressure Liquid Chromatography

High pressure liquid chromatography (HPLC) was used to measure concentrations of formaldehyde, acetaldehyde, and acrolein in the ame. This technique was selected because quantication of formaldehyde by GC/FID is not recommended[58]. Using HPLC for analysis of the aforementioned carbonyl compounds is mentioned in the literature[59, 60]. The procedure for analysis is outlined in the ASTM Standard Test Method for Determination of Formaldehyde and Other Carbonyl Compounds in Air[61]. The method involves drawing sample gases through a cartridge coated with 2,4-dinitrophenylhydrazine (DNPH) reagent. The carbonyl species form stable derivatives with the DNPH reagent, and are subsequently eluted from the cartridge. The derivatized liquid sample is analyzed for the parent carbonyls using HPLC. A schematic of the HPLC instrument is shown in Figure 3.14. The instruments consists of a reservoir of mobile phase, a pump, an injector, a separation column, and a detector. The mobile phase is a mixture of polar and non-polar solvents. It carries the injected sample through the stationary column and compounds are separated due to partitioning between the two phases. The partitioning behavior is a strong function of polarity, so separation of compounds with a range of polarities is obtained by changing the mobile phase mixture.

Figure 3.14: Schematic of HPLC Setup from [62]

Chapter 3. Experimental Apparatus and Analytical Methodology

54

The amount of time a given component spends in the separation column is called the retention time. The retention time of a given component remains the same provided the mobile phase, stationary phase, temperature control, and liquid owrates remain constant. As each component falls onto the detector a quantitative response in the form of a peak is generated. A series of peaks with the retention time on the x-axis and the detector ouptut (e.g. voltage) on the y-axis is called a chromatogram (see Appendix C for an example). The peak retention time is used to identify each compound, and the peak area is used to determine the quantity of the compound. The HPLC system used in this study was a Perkin Elmer series 200 quatenary LC pump coupled with a Perkin Elmer 785A UV/Vis detector. The instrument was located in the Department of Chemistry ANALEST Laboratory. Samples are injected by a Perkin Elmer auto-sampler through a Rheoudyne injector. The TurboChrome client/server software package was used to remotely control the HPLC equipment. Additional details of the HPLC technique are discussed in the following subsections.

DNPH Sampling Procedure Samples of carbonyl compounds were obtained by passing the sample gas through DNPH coated sampling cartridges23 . Two cartridges (i.e. main and backup) were connected, in series, downstream of the dual-stage pump, and the pump was started. The owrate of the sample was measured with a bubble meter downstream of the cartridges. The total amount of gas passing through the cartridges is the product of the owrate and the sampling time. The second cartridge (i.e. backup) was used to identify breakthrough of carbonyl compounds from the rst cartridge. A liquid sample was then obtained by elutriating the cartridge with 5 mL of 99% pure acetonitrile. The eluted sample was transferred to a 25 mL vial and stored at 4 C. Vials with septum caps were used to facilitate auto-sampling. The liquid solutions containing acetonitrile and carbonyl derivatives were then analyzed using the HPLC technique. Each DNPH cartridge is capable of holding approximately 75 mg of total carbonyls. Tests were conducted to determine the ideal gas sampling time required to prevent exceeding the cartridge capacity. The experiments involved lighting a methyl butanoate ame and varying the sampling duration. Samples were obtained from two points between the burner ports; one point is near the fuel port (2 mm from the fuel port) and one
23

SUPELCO LpDNPH S10 cartridges

Chapter 3. Experimental Apparatus and Analytical Methodology

55

point is near the ame front (7.5 mm form the fuel port). The point near the fuel port represents a region of high owrate and low concentration of carbonyls, while the point near the ame represents a point of low owrate and high concentration of carbonyls. The main and backup cartridges were then tested using the HPLC technique. The presence of species on the backup cartridge indicates a failure condition, with breakthrough of samples from the rst cartridge. Table 3.5 summarizes the results.
Distance from fuel port (mm)
2 2 2 2 2 7.5 7.5 7.5 7.5 7.5

Duration (min:sec)
2:42 3:22 5:02 7:12 10:30 2:02 3:45 4:33 6:29 10:12

Breakthrough (Yes/No)
No No No Yes Yes No No Yes Yes Yes

Table 3.5: Determination of ideal sampling times for carbonyl compounds using the DNPH technique

It was determined that the ideal sampling duration is between 2 and 3 minutes for all points in the ame. All subsequent experiments used this sampling duration. Backup cartridges were always used during gas sampling. They were then stored at 4 C, and a few were randomly selected for testing on the HPLC. Backup cartridges were also selected for HPLC testing if the mass of total carbonyl compounds on the corresponding main cartridge was greater than 60 mg. However, the failure condition was never observed for sampling times between 2 and 3 minutes. Following these procedures ensured accurate sampling of carbonyl compounds using the DNPH cartridges. HPLC Measurement Procedure Before sampling, the mobile phase solvents were degassed by bubbling helium for 5-6 minutes. The ow lines were then purged with water and acetonitrile, separately. After purging, the mobile phase ow was brought in-line with the separation column. A 0.5

Chapter 3. Experimental Apparatus and Analytical Methodology

56

m frit and a guard cartridge system24 were used to prevent damage to the separation column. The system was allowed to equilibrate by pumping 60% acetonitrile and 40% water at 1 mL/min for 20 minutes. The UV/Vis detector was operated at 360 nm with a range of 0.1. The experiments used an LC-18 reverse phase column25 that operates on the basis of hydrophilicity and lipophilicity. The stationary phase consists of an ocatdecyl ligand of covalently bonded n-alkyl chains in a silica based packing. The less polar the analyte, the longer its retention. The length, ID, and particle packing size are 25 cm, 4.6 mm, and 5 m, respectively. The system was calibrated before each set of experiments using standard carbonylDNPH26 mixes with known concentrations of various carbonyl compounds. The samples were then placed in the HPLC auto-sampler and ready for injection. The HPLC pump draws a set ratio of water and acetonitrile at 1.2 mL/min, as prescribed in the HPLC method. After injection of 20 L of sample, a gradient elution method was used to separate the components. Initially, the mobile phase mixture was 50% acetonitrile and 50% water. After 4 minutes, the pump increased the acetonitrile concentration to 100% over the next 25 minutes, via a linear gradient. The mobile phase ow then returned to the initial 50:50 ratio and was held there for an additional 5 minutes. The concentrations obtained from the HPLC/UV-Vis technique were for the liquid solutions consisting of acetonitrile and carbonyl-derivatives. The concentration of the parent carbonyl in the gas samples was obtained via a series of calculations. These calculations are shown in Appendix D.

3.5.4

Temperature Measurement

The ame temperature prole was obtained by measuring the local temperature at various regions in the ame. The measurements were obtained by the most direct method: inserting a thermocouple into the ame. The thermocouple was small compared to the thickness of the ame front so as to not disturb the ame. The primary drawbacks of using a thermocouple are the aerodynamic wake behind left the ame front and the possible catalytic activity of the thermocouple material[54]. The thermocouple measures the temperature by employing the dierence in ther24

Phenomenex SecurityGuardTM 25 SUPELCOSILTM LC-18, 5 m, 25 cm 26 SUPELCO Carbonyl-DNPH Mix 1

Chapter 3. Experimental Apparatus and Analytical Methodology

57

moelectrical properties of dierent metals. When two dissimilar conductors are welded together, an electric potential is generated which is proportional to the dierence in temperature conducted by each metal. Thermocouples made of thin noble metal wires such as platinum and rhodium are advantageous for the following reasons: they allow a high resolution to be obtained; the aerodynamic disturbance of the ame front is minimized; and the materials can withstand high temperature environments [54]. A schematic of the thermocouple apparatus is shown in Figure 3.15. The ame temperature was measured by inserting the apparatus between the two burners. The two R-type thermocouple wires27 (legs) are made of dissimilar metals butt-welded at a junction. The positive leg is made of pure platinum while the negative leg is comprised of 87% platinum and 13% rhodium. Each wires diameter is 254 m while the butt-welded junction is approximately 500 m. The wires are housed in ceramic tubes which provide support while spreading the wires apart. The junction, at which the temperature is measured, lies between the ceramic tubes. At the opposite end, the ceramic tubes are xed to an aluminum plate which is mounted on a sliding rail. A spring is used to pull the ceramic tubes together and keep the wires taught at the other end. The positive and negative legs are connected to an R-type extension wire which carries the measured signal to a digital thermometer28 .

Figure 3.15: Thermocouple Schematic

27 28

Omega R R-type butt-welded unsheathed ne-gauge thermocouple wires Digisense R DualLogR R Thermocouple Thermometer

Chapter 3. Experimental Apparatus and Analytical Methodology Correction for Radiation Losses

58

The temperature measured by the thermocouple diers from the true ame temperature due to aerodynamic, thermal, and/or chemical perturbations. The methods of minimizing the eect of these perturbations are discussed in detail by Fristrom and Westenberg [50]. However, even if all disturbances are minimzed, the thermocouple will register a dierent temperature than the true stream temperature due to radiation losses. Correcting for these losses is estimated by equating the heat transferred to the thermocouple from the gas to the heat lost by radiation from the wires. The equation given for a spherical device is: Tg Tc = where Tg is the true gas temperature (K) Tc is the measured gas temperature (K) is the emissivity of the thermocouple element (dimensionless)
W is the Stefan-Boltzmann Constant = 5.67x1008 ( m2 K 4 ) 4 d (Tc4 Tw ) 2k

(3.2)

d is the wire diameter (m)


W k is the thermal conductivity of gas ( mK )

Tw is the wall (ambient) temp. to which heat is radiated = 300 (K) The thermal conductivity, k, of the gases was estimated as the thermal conductivity of air at the measured temperature. The thermal conductivity of air at various temperatures was obtained from the CRC Handbook (CRC online). Linear interpolation and extrapolation was used to determine thermal conductivity at temperatures not listed in the handbook. The emissivity, , of the thermocouple element was obtained from the study by Bradley and Entwistle [63].

Chapter 4 Modeling
The experimental ame setup was modeled using the CHEMKIN 4.0 software package1 [37]. CHEMKIN is a set of software tools useful when dealing with chemical kinetic problems. The package contains a number of reactor models that can be used to represent the dierent real-world systems via a graphical user interface (GUI). Our opposed-ow diusion ame was modeled using the OPPDIF code. As previously mentioned, the opposed-ow diusion ame conguration has an axisymmetric geometry consisting of two concentric, circular ports facing each other. A simplied representation of the opposed-ow geometry is shown in Figure 4.1. Fuel is fed through the bottom port while oxidizer is fed through the top. The opposing ports produce an axisymmetric ow eld with a stagnation plane lying in the middle. A at diusion ame lying on the oxidizer side of the stagnation plane is generated, since most fuels require more air then fuel by mass. The two-dimensinal ow eld is mathematically simplied to one dimension by assuming uid properties are a function of axial distance only. The onedimensional model predicts the temperature and species proles between the two burner ports. The sensitivity of species concentration and temperature to chemical kinetic and thermodynamic data can also be calculated [64].

4.1

Modeling Procedure

Initially, the CHEMKIN 4.0 GUI was used to set up a diagram of the opposed-ow diusion ame with two inlet streams. The next step was to generate the linking les
1

Distributed by Reaction Design www.ReactionDesign.com

59

Chapter 4. Modeling

60

Figure 4.1: Geometry of opposed-ow diusion ame from [64] for the OPPDIF code. This required using the pre-processors to access three important information les: i.) the gas-phase kinetics le; ii.) the thermodynamic data le; and iii.) the gas phase transport data le. More information on these les is given below. The CHEMKIN Gas-Phase Interpreter reads the rst two les and generates the CHEMKIN linking le. The third le is used by the TRANSPORT Preprocessor to generate the Transport Linking File. Next, the characteristics of the reactor and inlet ows were inputed. This includes the velocity of each inlet stream, initial concentrations, pressure, burner conguration, temperature, and a number of solution method options. An input le was then created from these user-dened parameters, and the OPPDIF model was run. The OPPDIF code outputs a text le containing the solution. The Solution Export Utility was used to convert this text le into a comma separated values le format that was readable by Microsoft Excel.

4.2

Details of Input Files

The gas-phase kinetics le identies all the gaseous species present, and it provides a userdened chemical kinetic mechanism for the production and consumption of these species. The chemical kinetic mechanism details each reaction taking place and the appropriate

Chapter 4. Modeling

61

reaction rate parameters in the Arrhenius form, as shown in Equation 4.1. The gas phase kinetic le conforms to the CHEMKIN input format. kA = A (T /Tref )n exp where k is the reaction rate A is the Arrhenius constant T is temperature Tref is the reference temperature (usually 298 K or 25 C) n is the pre-exponential factor Ea is the activation energy R is the ideal gas constant The thermodynamic data le contains all the thermodynamic data for the species. Contained within this le are the species name, elemental composition, electronic charge, and phase. In addition, fourteen polynomials tting coecients are provided to calculate enthalpy and entropy at any temperature. The thermodynamic data le conforms to the NASA formatting requirements. The gas transport data le contains molecular parameters for each molecule. Included is this le are the Lennard-Jones well depth in Kelvin, the polarizibility in cubic angstroms, the rotational relaxation and collision number, the dipole moment in Debyes, the Lennard-Jones collision diameter in angstroms, and the geometrical conguration of the molecule. The gas transport data conforms to the CHEMKIN formatting requirements. This study uses the methyl butanoate chemical kinetic mechanism developed by Dr. Sandro Ga 2 . Dr. Sandro Ga also provided the gas transport data le. The thermol l dynamic data le was obtained from the website of the Lawrence Livermore National Laboratory Combustion Chemistry Group [65]. Ea RT (4.1)

University of Toronto Department of Mechanical and Industrial Engineering Combustion Research Group

Chapter 5 Results and Discussion


The primary goal of this study is to understand the eect of molecular structure (i.e. eect of unsaturation) on the combustion of FAMEs. This is accomplished by comparing opposed ow diusion ame species proles of methyl butanoate and methyl crotonate, which are surrogates for saturated and unsaturated long chain FAMEs, respectively. Any observed dierences between the ame species prole will be explained by oxidation reaction pathways analysis. The secondary goal of this study is to validate a chemical kinetic model for methyl butanoate, a short chain FAME used as a modeling surrogate for the long chain FAMEs found in biodiesel. The mechanism was developed by a team of individuals in our Combustion Research Group. The mechanism was validated against experimental data obtained from an opposed ow diusion ame, jet-stirred reactor, and plug ow reactor. The opposed ow diusion ame studies are the contribution of this thesis. The ame measurements are compared against modeling predictions obtained from Chemkin 4.0.

5.1

The Role of FAME Molecular Structure in Combustion

It was previously mentioned that FAME molecular structure can alter the emissions proles from engines operating on biodiesel. However, there is little understanding of the role of unsaturation (carbon-carbon double bonds) on the fundamental combustion properties of the fuel. The following sections present opposed ow diusion ame emissions proles generated by a saturated FAME (i.e. methyl butanoate) and an unsaturated 62

Chapter 5. Results and Discussion

63

FAME (i.e. methyl crotonate), and the dierences between the two proles are rationalized by oxidation pathway analysis. Before discussing the actual results, the general pathways for fuel consumption will be discussed.

5.1.1

Major Reaction Pathways for the Oxidation of Methyl Butanoate

The original methyl butanoate mechanism proposed by Fisher et al. [8] was analyzed to identify the initial reaction pathways responsible for fuel consumption. The mechanism identies hydrogen abstraction and decomposition as the key mechanisms by which the fuel is consumed. In Figure 5.1, it is evident that hydrogen can be abstracted from any of the carbon atoms to form one of four possible radical species. As previously mentioned, the abstraction of H occurs preferentially on secondary carbon atoms, followed by primary carbon atoms. Note that the pathways presented here are all the possible reactions. The most important pathways for fuel oxidation in the opposed ow diusion ame are presented later. At higher temperatures, the decomposition of methyl crotonate becomes increasingly important. The fuel consumption is driven either by unimolecular thermal decomposition or by decomposition via collision with a third body that causes fragmentation of the parent fuel. Figure 5.2 displays all the possible decomposition reactions for methyl butanoate. As previously mentioned, all these reactions are not necessarily important in the opposed ow diusion ame chemistry.

Chapter 5. Results and Discussion

64

methyl butanoate H2 C O C H2 C CH3 O -H -H H2 C O C H2 C CH3 O H3C H C O C H2 C CH3 O H3C H2 C C H2 O C CH2 O

H3C

-H

H2C

or H2 C O C H C CH3 O

H3C

Figure 5.1: Hydrogen abstraction from carbons atoms in methyl butanoate

methyl butanoate O C H2 C CH3 O H3C H2 C O C H2 C CH3 O H3C H2 C C H2 O C O

H2 C

+ CH3 O C H2 C CH3 O + C2H5 O C CH3 O + CH3O C3H7 H3C O C H2 + C

+ CH3

H2 C

Figure 5.2: Decomposition reaction pathways for methyl butanoate

Chapter 5. Results and Discussion

65

5.1.2

Major Reaction Pathways for the Oxidation of Methyl Crotonate

Currently there is no detailed chemical kinetic mechanism for the oxidation of methyl crotonate. Dr. Ga and coworkers, including the author of this paper, are in the process l of developing such a mechanism. The methyl butanoate mechanism proposed by Fisher et al. [8] contains some reactions for methyl crotonate, since methyl crotonate is a possible intermediate in methyl butanoate combustion. One pathway for forming methyl crotonate from methyl butanoate is by a series of hydrogen abstraction reactions, as shown in Figure 5.3. The major reaction pathways for the oxidation of methyl crotonate can be predicted from the aforementioned analysis of methyl butanoate oxidation. At low and intermediate temperatures methyl butanoate is primarily consumed by hydrogen abstraction from the carbon atoms. Therefore in the same temperature range, hydrogen abstraction is the primary method of consuming methyl crotonate, as shown in Figure 5.4. Hydrogen abstraction can occur on any of the carbon atoms in the fatty acid portion of the molecule. This is because the radicals A, B, and C in the gure are in equilibrium due to internal isomerization. The radicals C and D are equivalent due to resonance. The decomposition of methyl crotonate is an important consumption pathway at higher temperatures. Figure 5.5 shows the possible C-O and C-C bonds that can be broken to form radical species. These decomposition pathways are more signicant at high temperatures. In the oxidation of alkenes, the biradical O was shown to attack the C=C double bond to form intermediate species. The double bond in methyl crotonate is susceptible to the same attack, as shown in Figure 5.6. The biradical O reacts with methyl crotonate to form aldehyde radicals and simpler ester radicals.

Chapter 5. Results and Discussion

66

methyl butanoate O H2 C H 3C C H2 C O CH3 +

radical O X H3C H C C H2 C O CH3 + XH

radical O H C H 3C C H2 C O CH3 + X H3C

methyl crotonate O H C C H C O CH3 + XH

Figure 5.3: Reaction pathway for forming methyl crotonate from methyl butanoate

H2C

H C

O C C CH3 O
H3C
resonance

methyl crotonate O H C C C O H

CH3 -H

-H O -H CH3 O H C H3C C or C H3C O C C H O CH3 O C O CH3

O H C H3C C H C O CH2

H C H2C C H

C
internal isomerization

Figure 5.4: Hydrogen abstraction from carbons atoms in methyl crotonate

Chapter 5. Results and Discussion

67

O H C C H + CH3 C O CH3 H3C

methyl crotonate O H C C C O H

O CH3 H3C H C C H + CH3 C O

O C O + C H3C C H CH3 H3C

O H C C H + CH3 O C

Figure 5.5: Decomposition reaction pathways for methyl crotonate

H C H3C C H

O O + C O +O O H C C C H O CH3 CH3

methyl crotonate

H3C

+O O C H3C O + C H2C O CH3

Figure 5.6: Reactions with biradical O for methyl crotonate

Chapter 5. Results and Discussion

68

5.1.3

Comparison of Opposed Flow Diusion Flame Emissions Proles

Methyl crotonate and methyl butanoate were used to generate opposed ow diusion ames as described in Section 3. Gas samples were withdrawn from a series of locations by moving the sampling probe from the fuel port to the oxidation port in a stepwise manner. The luminous blue ame front was located at approximately 8.5 mm above the fuel port. The temperature and emission proles of methyl butanoate and methyl crotonate are now compared to determine the role of FAME molecular structure (i.e. the double bond in methyl crotonate) in combustion. Figure 5.7 shows a simplied diagram of the burner setup rotated clockwise by 90 degrees. The species concentration and temperature proles in the following sections are oriented with respect to this diagram; the distance from the fuel port is plotted on the X-axis, while the measured parameter is plotted on the Y-axis. Figures 5.8 to 5.14 display the proles for the two fuels.

Figure 5.7: Diagram of the burner setup to clarify the orientation of experimental proles

Chapter 5. Results and Discussion

69

2000 1800 1600 TEMPERATURE (K) 1400 1200 1000 800 600 400 200 0 0 5 10 15 DISTANCE FROM FUEL PORT (mm) 20

TEMP- MC TEMP- MB

Figure 5.8: Measured temperature proles in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames

12

10 MOL FRACTION (%)

CO- MB CO2- MB CO- MC CO2- MC MB MC

0 0 5 10 15 DISTANCE FROM FUEL PORT (mm) 20

Figure 5.9: Measured concentration proles for fuel (MB or MC), CO, and CO2 in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames

Chapter 5. Results and Discussion


6000 C2H4- MB C2H4- MC C2H2- MB C2H2- MC CH4- MB CH4- MC

70

5000 MOL FRACTION (PPM)

4000

3000

2000

1000

0 2 3 4 5 6 7 8 DISTANCE FROM FUEL PORT, mm 9 10

Figure 5.10: Measured concentration proles for CH4 , C2 H4 , and C2 H2 in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames

2000 1800 1600 MOL FRACTION (PPM) 1400 1200 1000 800 600 400 200 0 2 4 6 8 DISTANCE FROM FUEL PORT (mm) 10 C3H6- MB C2H6- MB C2H6- MC C3H6- MC CH3OH- MC

Figure 5.11: Measured concentration proles of C3 H6 and C2 H6 in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames, and CH3 OH in the methyl crotonate ame

Chapter 5. Results and Discussion


70 60 MOL FRACTION (PPM) 50 40 30 20 10 0 2 3 4 5 6 7 8 DISTANCE FROM FUEL PORT, mm 9 10 2-C4H6- MB 2-C4H6- MC C3H8- MB C3H8- MC CH3CHO- MB CH3CHO- MC

71

Figure 5.12: Measured concentration proles of 2-C4 H6 , C3 H8 , and CH3 CHO in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames

600 1-C4H8- MB 1-C4H8- MC C3H4- MB C3H4- MC 1,3-C4H6- MB 1,3-C4H6- MC CH2O- MC CH2O- MB

500 MOL FRACTION (PPM)

400

300

200

100

0 2 3 4 5 6 7 8 DISTANCE FROM FUEL PORT, mm 9 10

Figure 5.13: Measured concentration proles of C4 H8 , C3 H4 , 1,3-C4 H6 , and CH2 O in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames

Chapter 5. Results and Discussion


70 60 MOL FRACTION (PPM) 50 40 30 20 10 0 2 3 4 5 6 7 8 DISTANCE FROM FUEL PORT, mm 9 10

72

C5H12- MB C5H12- MC C6H6- MC C3H4O- MC C3H6O- MC C4H10- MB C4H10- MC

Figure 5.14: Measured concentration proles of C5 H12 and C4 H10 in the methyl butanoate (MB - closed symbols with lines) and methyl crotonate (MC - open symbols without lines) ames, and C6 H6 , C3 H4 O, and C3 H6 O in the methyl crotonate ame

Table 5.1 shows the peak temperatures obtained while burning methyl butanoate and methyl crotonate are 1758 K and 1750 K, respectively. In addition, the ame temperature proles for the two fuels are very similar (Figure 5.8). The similarity in measured temperatures is expected because the fuels have similar heating values. The heating value is not available for the two fuels, but Demirbas [66] has shown that a fuels heating value can be directly correlated to the density and viscosity of the fuel. Since the two fuels are structurally similar, the physical chemical properties, and therefore, the heating values would be comparable. However, methyl crotonate would have slightly lower heating value because the C=C double bond lowers the heat content of the fuel [66]. Concentration proles were obtained for a number of hydrocarbon and oxygenated species (Figures 5.9 to 5.14). The fuel reactivity, CO, CO2 , CH4 , C2 H6 , CH2 O, and CH3 CHO proles for the two fuels are very similar. The methyl crotonate ames shows consistently higher levels of C2 H2 , C3 H4 , C3 H4 , 1-C4 H8 , 2-C4 H6 , C4 H10 , C5 H12 , 1,3C4 H6 . In addition, methyl crotonate proles are available for CH3 OH, C3 H4 O, C3 H6 O, and C6 H6 , but the same species were not detected in methyl butanoate. Consistently

Chapter 5. Results and Discussion lower levels of C2 H4 and C3 H8 were detected in the methyl crotonate ame1 .

73

Table 5.1 summarizes the measured peak mole fractions in the two ames. Methyl butanoate and methyl crotonate produce similar peak mole fractions for carbon monoxide (CO), carbon dioxide (CO2 ), methane (CH4 ), ethane (C2 H6 ), formaldehyde (CH2 O), and acetaldehyde (CH3 CHO). However, the methyl crotonate ame produces approximately twice the mole fraction of propene (C3 H6 ) and acetylene (C2 H2 ); half the mole fraction of ethylene (C2 H4 ) and propane (C3 H8 ); ve to seven times more propyne (C3 H4 ), 1-butene (1-C4 H8 ), 2-butyne (2-C4 H6 ), butane (C4 H10 ), and n-pentane (C5 H12 ); and twelve times more 1,3-butadiene (1,3-C4 H6 ). In addition, the methyl crotonate ame produces methanol (CH3 OH), acrolein (C3 H4 O), acetone (C3 H6 O), and benzene (C6 H6 ) in detectable quantities, while these species were below the detection limit in methyl butanoate ames.
Opposed Flow Diffusion Flame Measured Parameter Temperature (K) Carbon Dioxide CO2 (%) Carbon Monoxide CO (%) Methane CH4 (ppm) Formaldehyde CH2O (ppm) Ethane C2H6 (ppm) Ethylene C2H4 (ppm) Acetylene C2H2 (ppm) Propene C3H6 (ppm) Propyne C3H4 (ppm) Acetaldehyde CH3CHO (ppm) 1-Butene 1-C4H8 (ppm) 1,3-Butadiene 1,3-C4H6 (ppm) Acrolein C3H4O (ppm) Propane C3H8 (ppm) n-pentane C5H12 (ppm) 2-Butyne 2-C4H6 (ppm) Butane C4H10 (ppm) Acetone C3H6O (ppm) Benzene C6H6 (ppm) Methanol CH3OH (ppm) Experimental Results (MC) 1758 10 3 2185 280 1802 2567 4781 1394 434 40 353 568 50 31 13 34 61 21 16 617 Experimental Results (MB) 1750 10.1 3 1957 222 1405 4842 1951 790 62 43 58 48 <3 62 2 6 12 <3 <5 <5 Ratio (MC/MB) 1.0 1.0 1.0 1.1 1.3 1.3 0.5 2.5 1.8 7.0 0.9 6.1 11.8 0.5 6.5 5.7 5.1

Table 5.1: A comparison of measured peak species mole fractions in methyl butanoate (MB) and methyl crotonate (MC) ames

See Appendix B for structures of measured chemical species

Chapter 5. Results and Discussion

74

5.1.4

Rationale for Dierences in Methyl Butanoate and Methyl Crotonate Emissions Proles

The aforementioned experimental results indicate noticeably dierent levels of many species. In general, the unsaturated methyl crotonate leads to the formation of more unsaturated species. Analogous reaction pathways for the oxidation of methyl butanoate and methyl crotonate can be used to explain these observed dierences. Decomposition and hydrogen abstraction reactions were shown to be the main pathways during combustion. These reactions can lead to the production of C3 H7 and C3 H5 from methyl butanoate and methyl crotonate, respectively. Figure 5.15 shows the decomposition (#1, #3, #4) and hydrogen abstraction (#2) reactions responsible for forming the n-propyl radical (C3 H7 ) from methyl butanoate. C3 H7 further reacts to form propane (C3 H8 ) and ethylene (C2 H4 ) by H addition (#6) and -scission (#5), respectively. This explains why concentrations of these species are greater in the methyl butanoate ame. The analogous decomposition (#7, #10, #11) and hydrogen abstraction (#8) reactions for methyl crotonate lead to the formation of the 1-propenyl radical (C3 H5 ), as shown in Figure 5.15. In addition, methyl crotonate can decompose (#9) directly to C3 H5 and the methoxycarbonyl radical (CH3 OCO) by cleavage between hydrocarbon chain and the methyl ester group. The possibility of this cleavage is greater in methyl crotonate than methyl butanoate due to the greater electronegativity of the C=C double bond near the ester group. The double bond attracts electrons away from the adjacent single bond, thereby making it weaker and easier to break. C3 H5 then reacts to form propene (C3 H6 ) by H addition (#14), and propyne (C3 H4 ) and acetylene (C2 H2 ) by -scission (#12a, #12b). C2 H2 then undergoes pyrolysis (#15) in the presence of smaller hydrocarbon to form 1-butene (1-C4 H8 ), 2-butyne (2-C4 H6 ), 1,3-butadiene (1,3-C4 H6 ) and benzene (C6 H6 ). Figure 5.15 also shows that CH3 OCO decomposes (#16) to CH3 O and CO, and then CH3 O undergoes H addition (#17) to form CH3 OH. This explains the higher concentrations of all the aforementioned stable species in the methyl crotonate. Acrolein (2-propenal C3 H4 O) was measured in the methyl crotonate ame but was not detectable in the methyl butanoate ame. Figure 5.15 shows that acrolein can be derived from methyl crotonate by a series of decomposition and H addition reactions (#18, #19, #20, #22). In the case of methyl butanoate, it is evident that the analogous

Chapter 5. Results and Discussion

75

pathways (#22, #23, #24, #25) would lead to propenal (C3 H6 O), instead of acrolein. Propanal concentrations were below the detection limit for both fuels. Benzene (C6 H6 ) was measured in the methyl crotonate ame but not in the methyl butanoate ame. This can be attributed to the higher levels of acetylene, propyne, and C4 hydrocarbons present in the methyl crotonate ame. It was previously shown that, under richer conditions, these species can form complex aromatic compounds, PAHs, and ultimately soot. The double bond in methyl crotonate is responsible for the higher levels of acetylene, propyne, 1,3-butadiene, 1-butene, 2-butyne, and benzene. It can be inferred that unsaturated long chain FAMEs will have a similar tendency to create the aforementioned soot precursors in greater concentrations than saturated long chain FAMEs. Therefore, an important observation of this study is that unsaturated FAMEs have a greater tendency to form soot precursors.

Chapter 5. Results and Discussion

76

O H3C C H2 propanal CH H2C C H +H O H2 C C H2 CH - CH3 H C H3C +H H C H3C H C

O CH acrolein (2-propenal)

#21
O C H CH

#25

+H

#20
O C H CH

- CH3 H2 C H3C +H H2 C H3C

#24
O C H2 CH

#19
O C H C

#23
O C H2 C

#18
- CH3O O Methyl Crotonate H3C CH3 O +X - CH3

- CH3O O H C C H C O +X CH3

#22

Methyl Butanoate H3C - CH3 O H2 C H3C C H2 C O

H2 C

- CH3OCO

C H2

#7
- XH O H2 C C O CH2 - CO2 H3C O H C C H C O

#9

#8

- XH O H C C C H O CH2

#1

#2

H3C

H3C

C H2

#10
H C H3C

#11
- CO, - CH2O

- CO2

#3
H2 C H3C

#4

- CO, - CH2O

propenyl radical

C H +H

n-propyl radical

C H2

H +H propyne H3C H2 C + C CH

#12a
sion

#14
H C H3C CH2 propene O C O CH3 - CO CH3 + H

#5
CH2

#6

CH3

H2C

or
CH3 CH3 + HC CH

ethylene

H3C

#12b
acetylene

!-sc is

!-scission

propane

#16
O

#15
1-C4H8 1,3 - C4H6 2-C4H6 C6H6

pyrolysis

#17
1-butene 1,3 - butadiene 2-butyne benzene

HO CH3 methanol

Figure 5.15: Analogous reaction pathways for methyl butanoate and methyl crotonate oxidation in the opposed ow diusion ame

Chapter 5. Results and Discussion

77

5.2

Chemical Kinetic Modeling of Methyl Butanoate

The detailed chemical kinetic mechanism used in these calculations was developed by Dr. Ga and coworkers, one of them being the author of this thesis [10]. The original l Fisher et al. mechanism [8] was modied to better reproduce jet-stirred reactor data obtained at the CNRS, Laboratoire de Combustion et Syst`mes Ractifs. The modied e e mechanism was then tested against experimental data from a plug ow reactor and opposed ow diusion ame; the latter being the contribution of this thesis study. A C4 sub-mechanism was also added to the overall mechanism to model the formation of C4 hydrocarbon species. The revised mechanism [10] consists of 1498 reversible reactions involving 295 species. Several reaction rate parameters from the original mechanism [8] were modied, as shown in Table 5.2. Some of these reaction rate modications were performed by the author, as part of this thesis study.

Reaction

Fisher et al. rate constants (A / n / Ea)

Modied rate constants (A / n / Ea)

CH3CH2CH2(C=O)OCH3 + H

CH3CH2CH(C=O)OCH3 + H2

2.52E+14 /

0.00

300

1.00E+14 /

0.00

300

CH3O (+M)

CH2O +

H (+M)

5.45E+13 /

0.00 / 13500 1.38E+21 / -6.65

/ 33190

C2H3 + O2

CH2O +

HCO

1.70E+29 /

-5.31 / 6500

1.66E+13 / -1.39 / 1013

CO + CH3O

CH3OCO

1.50E+11 /

/ 3000

1.50E+6 /

2.02

/ 5730

CO2 + CH3

CH3OCO

1.50E+11 /

/ 36730 4.76E+7 /

1.54

/ 34700

Table 5.2: Modied reaction rate constants [10, 8, 43].

Chapter 5. Results and Discussion

78

5.2.1

Modication of Reaction Rates - Contributions of this Study

A chemical kinetic study on the oxidation of dimethyl carbonate [43] was previously discussed in Section 2.5.3. The study by Glaude and coworkers used quantum mechanical estimates to provided new reaction rate constants for decomposition reactions of CH3 OC =O. The decomposition reactions and their respective rate constants are shown in Figure 5.3.
O CO2 + CH3 A = 4.76 E 7 n = 1.54 Ea = 34,700 C CH3 O A = 1.55 E 6 n = 2.02 Ea = 5,730 CO + CH3O

Table 5.3: Decomposition pathways of CH3 OC =O and corresponding reaction rate constants [43].

The author of this thesis was responsible for the testing and implementation of these rate constants in the chemical kinetic mechanism for methyl butanoate. These modied reaction rate constants were not included in the mechanism published by Ga and l coworkers [10]. The following section displays plots of the model predicted values and the experimental data. Modeling results are shown for the published Ga et al. mechanism l [10] (solid lines, old), as well as the revised model, which includes the aforementioned reaction rate constants from Glaude et al. [43] (dashed lines, new).

5.2.2

Model Validation with Opposed Flow Diusion Flame Results

The methyl butanoate opposed ow diusion ame was generated as described above. Gas samples were withdrawn from a series of locations by moving the sampling probe from the fuel port to the oxidation port in a stepwise manner. The luminous blue ame front was located at approximately 8.5 mm above the fuel port. The experimental data was compared against model generated values using Chemkin 4.0 running on the methyl butanoate mechanism. Molecular species concentration proles

Chapter 5. Results and Discussion

79

were obtained from the oxidation of methyl butanoate in an opposed ow diusion ame for the following species: methyl butanoate, CO, CO2 , CH4 , C2 H4 , C2 H2 , C3 H6 , C2 H6 , CH2 O, C4 H8 , C3 H8 , pC3 H4 , CH3 CHO, and 1,3-C4 H6 2 . Figures 5.16 to 5.19 show the experimental and modeling results for the aforementioned species. The published Ga et l al. model [10] is shown in solid lines and marked old, while model with rate constants from the study by Glaude et al. [43] is shown as dashed lines and marked new.
12 CO 10 MOL FRACTION (%) CO2 MB
old old old new new new

6 8 10 12 14 DISTANCE FROM FUEL PORT (mm)

16

18

Figure 5.16: Modeling predictions (small symbols with lines) and experimental results (large symbols without lines) for methyl butanoate (MB), CO, and CO2

See Appendix B for structures of measured chemical species

Chapter 5. Results and Discussion


6000 CH4 5000 MOL FRACTION (PPM) C2H4 C2H2
old old old new new new

80

4000

3000

2000

1000

0 2 4 6 8 DISTANCE FROM FUEL PORT (mm) 10

Figure 5.17: Modeling predictions (small symbols with lines) and experimental results (large symbols without lines) for CH4 , C2 H4 , and C2 H2

1800 C3H6 1600 C2H6 MOL FRACTION (PPM) 1400 1200 1000 800 600 400 200 0 2 4 6 8 DISTANCE FROM FUEL PORT (mm) 10 CH2O
old old new new old new

Figure 5.18: Modeling predictions (small symbols with lines) and experimental results (large symbols without lines) for C3 H6 , C2 H6 , and CH2 O

Chapter 5. Results and Discussion

81

200 C4H8 180 160 MOL FRACTION (PPM) 140 120 100 80 60 40 20 0 2 4 6 8 DISTANCE FROM FUEL PORT (mm) 10 C3H8 pC3H4 CH3CHO C4H6
old old old old old new new new new new

Figure 5.19: Modeling predictions (small symbols with lines) and experimental results (large symbols without lines) for C4 H8 , C3 H8 , pC3 H4 , CH3 CHO, and 1,3-C4 H6

Species Name (measurement units) Carbon Dioxide CO2 (%) Carbon Monoxide CO (%) Ethylene C2H4 (ppm) Methane CH4 (ppm) Propane C3H8 (ppm) Ethane C2H6 (ppm) Propene C3H6 (ppm) Propyne C3H4 (ppm) Acetylene C2H2 (ppm) 1-Butene 1-C4H8 (ppm) Formaldehyde CH2O (ppm) 1,3-Butadiene 1,3-C4H6 (ppm) Acetaldehyde CH3CHO (ppm)

Experimental Results 10.1 3.0 4842 1957 62 1405 790 62 1952 58 222 48 43

Model by Gail et al. 10.6 3.4 4841 2578 49 1077 1630 100 4572 187 866 22 22

Ratio (Gail/Exp) 1.05 1.13 1.00 1.32 0.79 0.77 2.06 1.61 2.34 3.22 3.90 0.46 0.51

Model in this Thesis 10.5 3.0 5257 3114 51 1442 1630 128 4777 180 593 24 19

Ratio (Thesis/Exp) 1.04 1.00 1.08 1.59 0.82 1.03 2.06 2.06 2.44 3.10 2.67 0.5 0.44

Model by Fisher et al. 10.6 3.6 4089 2590 49 1295 1452 108 4321 --1658 --25

Ratio (Fisher/Exp) 1.05 1.20 0.84 1.32 0.79 0.92 1.84 1.74 2.21 --7.47 --0.58

Table 5.4: A comparison of peak species mole fractions from experimental data and model-predicted values

Chapter 5. Results and Discussion

82

The results show fairly good agreement between the model predicted values and the experimental data. However, the consumption of methyl butanoate is under-predicted by the model (Figure 5.16). This is evident by the spatial delay in methyl butanoate consumption displayed by the model generated prole, when compared to the experimental data. The maximum species concentrations predicted by the model are shifted away from the fuel port (Figures 5.17 to 5.19). This is due to the under-prediction of methyl butanoate consumption; since the fuel consumption is delayed, the production of species is also delayed. It should be noted that the reactivity of methyl butanoate in the Fisher et al. model [8] was improved upon by modifying the Arrhenius rate constant of the hydrogen abstraction reaction forming the 2-methyl butanoate radical, as shown previously in Table 5.2. The modication of CH3 OCO decomposition rate constants further improved the methyl butanoate reactivity, as is evident by comparing the old and new curves. Table 5.4 compares the peak concentrations from the experimental data to the modelpredicted values. The peak concentrations of carbon monoxide (CO), carbon dioxide (CO2 ), methane (CH4 ), ethylene (C2 H4 ), propane (C3 H8 ), and ethane (C2 H6 ) are well predicted by the model. However, the model over-predicts the peak concentration of acetylene (C2 H2 ), 1-butene (C4 H8 ), propene (C3 H6 ), propyne (pC3 H4 ), and formaldehyde (CH2 O) by 2-4 times. The model under-predicts the peak concentrations of 1,3-butadiene (1,3-C4 H6 ) and acetaldehyde (CH3 CHO) by approximately one-half. 3 . l A comparison of the models by Fisher et al [8], Ga et al. [10], and this thesis study is also presented in Table 5.4. The gray-highlighted rows indicate which model produces better agreement with the experimental data. The three models perform similarly, with this studys model being marginally better for the majority of species. This studys model performs much better for formaldehyde due to the modication of two sets of reaction rate constants involving formaldehyde and the changes to CH3 OCO decomposition reaction rates, as shown in Table 5.2. The Ga et al. model enables the prediction of C4 l hydrocarbons, while the Fisher model does not. This is due to the C4 sub-mechanism included in the mechanism by Ga l. The modications to the CH3 OCO decomposition reaction rates have interesting effects on the mechanism. First, the rate of production of CO and CH3 O is reduced, thereby reducing the CO and CH3 O concentrations in the ame. This, in turn, reduces
3

See Appendix B for structures of measured chemical species

Chapter 5. Results and Discussion

83

CH2 O concentrations since CH2 O is formed by hydrogen abstraction from CH3 O. Many species concentrations were changed by this modication; thus, indicating the sensitivity of the mechanism to these reaction rate parameters.

5.2.3

Error Analysis of Major Dierences in Modeling and Experimental Results

The previous section showed that the model over-predicts the peak concentration of acetylene (C2 H2 ), propene (C3 H6 ), propyne (pC3 H4 ), formaldehyde (CH2 O), and 1-butene (C4 H8 ) by 2-4 times. The model under-predicts the peak concentrations of 1,3-butadiene (1,3-C4 H6 ) and acetaldehyde (CH3 CHO) by approximately one-half. These dierences can be attributed to either experimental error or inaccuracies in the mechanisms reaction rate coecients. Experimental Errors The GC/FID sampling method for hydrocarbons is unlikely to be a source of error because the results show good agreement between modeling and experimental results for methane, ethylene, ethane, and propane. The presence of an error in the sampling of hydrocarbon species would be evident in all the species proles; however, this is not the case. In addition, extra precautions were taken to ensure that the sampling systems credibility was not compromised, such as: i.) detecting and minimizing leakage into the sampling line, ii.) freezing reactions at the probe tip by creating a vacuum pressure of 710 - 730 mm Hg, iii.) minimizing aerodynamic disturbances by using the smallest possible sampling probe, iv.) passing calibration gases through the burner ports and sampling them before every experiment, and v.) adequately heating all the sampling lines to prevent condensation of sample gases. The accuracy of the measured concentrations of carbonyl compounds, such as formaldehyde and acetaldehyde, may have been compromised by the sampling method. These compounds were measured by passing the samples through cartridges coated with DNPH. The carbonyl compounds form stable derivatives with the DNPH reagent, are are subsequently eluted from the cartridge. The derivatized liquid samples were analyzed for their parent carbonyls using an HPLC/UV-vis setup in another laboratory. This sampling method required signicant human interaction to prepare and transport the samples between laboratories, thereby creating possible sources of error. The precautions mentioned

Chapter 5. Results and Discussion

84

above were followed when sampling these compounds, except the calibration method was dierent. Calibration for these compounds was not as thorough as the calibration of hydrocarbons because calibration gases were not available for formaldehyde and acetaldehyde. Therefore, it was not possible to pass a calibration gas through the burner ports and test the accuracy of this sampling method. Instead, only the HPLC/UV-vis instrument was calibrated using known concentrations of DNPH derivatives. It is recommended that future experiments either obtain calibration gases or use a GC/FID coupled with a methanizer to measure these compounds. The results from the GC/FID can then be compared to those obtained by the DNPH sampling method.

Inaccuracies in Model Reaction Rate Coecients Accurate reaction rates coecients are arguably the most important input into a chemical kinetic model. Such values are obtained by either experimental or theoretical methods, and they can be found in the literature. Of these, experimental methods are more accurate, but more dicult to obtain. Theoretical reaction rate coecients for a given reaction may vary signicantly depending the correlations used for their calculation. In addition, reaction rate coecient in chemical kinetic model are often times guessed. The guessing of these coecients is performed by selecting a value that best reproduces a given set of experimental data. Thus, there is a possible source of error present in the mechanisms reaction rate coecients. In order to identify possible sources of error, an A-factor sensitivity analysis and a rate of production analysis were performed for CH2 O, C2 H2 , pC3 H4 , C4 H8 , and C3 H6 . An A-factor sensitivity analysis provides rst-order sensitivity coecients for the species. The rst-order sensitivity coecient is dened as the normalized derivative of the species concentration with respect to the A-factor of an individual reaction. The rate of production analysis provides complementary information on the contributions of individual reactions on the net production rate of a species [64]. Both these analyses were conducted at the point of maximum concentration of each species. The reactions with highest rst order sensitivity coecients and rate of production values are the most likely to impact the predicted concentration of a species. Therefore, the data from the two aforementioned methods were analyzed to select the reactions which would have the highest eect on the species of interest. Table 5.5 shows the reactions that are most likely to have an impact of the maximum concentration of a particular

Chapter 5. Results and Discussion

85

species. The table also compares the reaction rate coecients given in the present model to those found on the National Institute of Standards and Technology (NIST) Scientic and Technical Databases - Chemical Kinetics [67]. Table 5.5 compares reaction rate values given in the present model [10] to those available in literature [67]. The rst column shows the species sensitivity and the correlation between species concentration and reaction rate; a plus (+) sign indicates a positive correlation, and increasing the reaction rate will increase the species concentration, while a minus (-) sign indicates a negative correlation, and increasing the reaction rate will decrease the species concentration. The second column gives the reaction of interest. The third and fourth columns show the Arrhenius equation constants (i.e. A, n, and Ea ) and the calculated reaction rate (cm6 mol1 s1 ) at the temperature corresponding with maximum species concentration. The rst row shows a reaction (#1) in which C2 H2 reacts with an H radical to form C2 H3 . The model calculated reaction rate for this reaction is less than that determined by Tsang and Hampson [68]. Increasing this reaction rate would increase the forward reaction rate, thereby decreasing the C2 H2 concentration. The second row shows a reaction (#2) in which C2 H2 reacts with a methyl radical (CH3 ) to form pC3 H4 and an H radical. Again, the model calculated reaction rate is less than the reaction rate determined in a study by Davis and coworkers [69]. Increasing this reaction rate would increase the conversion of C2 H2 to pC3 H4 . This would reduce the predicted C2 H2 concentration, but simultaneously increase the predicted pC3 H4 concentration. The eects of this trade-o can only be quantied by testing in the model. The reaction (#3) in the third row displays C3 H6 decomposing to C2 H3 and CH3 . The reaction rate provided in the model is less than that determined in a study by Hidaka and coworkers [70]. Reducing the reaction rate to the literature values would decrease the forward reaction rate, thereby increasing the concentration of C3 H6 . This is not desired since C3 H6 is already over-predicted by the model. However, decreasing the forward reaction rate would cause C2 H3 to decrease. Lower amounts of C2 H3 in the gas phase would drive reaction #1 forward due to Le Chateliers principle, thereby decreasing C2 H2 levels. Again, the trade-o between an increase in C3 H6 and a decrease in C2 H2 concentration would have to be quantied by running the model with the new values. The nal row in the table shows a reaction (#4) involving the iC3 H7 radical undergoing H abstraction to form C3 H6 and H2 . The sensitivity analysis revealed that CH2 O, C2 H2 , pC3 H4 , C4 H8 , and C3 H6 are all sensitive to this reaction. This reaction has not

Chapter 5. Results and Discussion

86

been studied in the past, so it is likely that the authors of the original methyl butanoate mechanism [8] guessed the reaction rate coecients. This leads to the possibility of one set of reaction rate parameters causing the over-prediction of a number of species. It is recommended that lower reaction rate values be tested to quantify the eects on decreasing CH2 O, C2 H2 , C4 H8 , and C3 H6 concentrations. However, care must be taken not to increase pC3 H4 concentrations by too much, since pC3 H4 has a negative correlation with the reaction rate. The aforementioned changes to reaction rates were attempted in this study. However, due to computational limitations, the eects of the changes could not be quantied. Any changes made to the existing model would not lead to a converging solution. The problem was identied to be the C4 sub-mechanism consisting of an additional 279 reactions, which causes a signicant increase in computational load. It is recommended that the model be reduced to minimize the computations, and then the eects of changes to specic reaction rate constants can be determined.
Species Sensitivity (+/-) Present reaction rate (A / n / Ea) reaction rate (cm6.mol-2.s-1) calculated at @ T (K) NIST reaction rate (A / n / Ea) reaction rate (cm6.mol-2.s-1) calculated at @ T (K)

Reaction

#1 C2H2 (-) C2H2 + H (+M) = C2H3 (+M)

3.11E+11 / 0.6 / 2589 2.44E+13 @1400 K

3.81E+40 / -7.27 / 7210 4.16E+17 @1400 K

C2H2 (-) pC3H4 (+)

#2 C 2H 2 + CH3 = pC3H4 + H

1.21E+17 / -1.2 / 16680 2.44E+13 @1200 K

1.35E+12 / 1.1 / -27110 3.29E+15 @1200 K

C3H6 (-) C2H2 (-)

#3 C3H6 = C2H3 + CH3

2.73E+62 / -13.3 / 123200 3.05E+21 @1200 K

8.00E+14 / 0 / 88030 8.01E+14 @1200 K

C4H8 (+) CH2O (+) C2H2 (+) C3H6 (+) pC3H4 (-)

#4 iC3H7 + H = C3H6 + H2

2.00E+12 / 0 / 0 2.00E+13 @1200 K

NONE

Table 5.5: A comparison of reaction rates given in the present model [10] to those available in literature [67]

Chapter 5. Results and Discussion

87

5.2.4

Major Reaction Pathways for the Oxidation of Methyl Butanoate in the Opposed Flow Diusion Flame

The major reaction pathways for methyl butanoate oxidation in the opposed ow diusion ame were determined via rate of production and consumption analysis. This involved studying the Chemkin output le to determine which reactions had the highest rates for methyl butanoate consumption. The rates of consumption of subsequent species was also studied until typical combustion intermediates were formed, such as CH3 , CH2 O, CH3 O, etc.. The rate of production analysis was conducted at a distance of 8.3 mm from the fuel port, which corresponds to a ame temperature of 1220 K. It should be noted that this temperature is lower than the maximum ame temperature of 1750 K. Figure 5.20 exhibits the major reaction pathways for methyl butanoate oxidation in the opposed ow diusion ame. The red arrows signify the most important reaction pathways. The weight of the blue arrows signies the relative rate of reaction, i.e., reactions with heavy arrows are more important that reactions with lighter arrows. The diagram displays reaction pathways for species until the point where the functional methyl ester group is consumed. Initially, the fuel is consumed primarily by H-abstraction reactions by H atoms due to the intermediate temperature range. The reactions proceed to form one of four C5 H9 O2 radical species. Of these, the formation of 2-methyl butanoate (CH3 CH2 CH (C=O)OCH3 ) and 3-methyl butanoate (CH3 CH CH2 (C=O)OCH3 ) radicals are favored because secondary C-H bonds are weaker than primary C-H bonds. The 2-methyl butanoate radical undergoes -scission to remove a methyl radical (CH3 ) and form methyl propenoate (CH2 CH(C=O)OCH3 ). Methyl propenoate then undergoes hydrogen addition to create formaldehyde (CH2 O) and propenoyl radical (C2 H3 CO). The 3-methyl butanoate radical undergoes thermal decomposition to remove propene (C3 H6 ) and form a methoxycarbonyl radical ( (C=O)OCH3 ). Subsequently, the methoxycarbonyl radical decays to form either a methoxy radical (CH3 O) and carbon monoxide (CO) or a methyl radical (CH3 ) and carbon dioxide (CO2 ). Of these two reactions, the one forming CO2 was found to be more signicant than the one producing CO. From a soot reduction standpoint this is not an ecient use of fuel-bound oxygen because two oxygen atoms are bonded to one carbon atom thereby wasting an oxygen atom in the

Chapter 5. Results and Discussion

88

functional methyl ester group. This conclusion is consistent with the ndings of other researchers [43, 44]. Alternative pathways for methyl butanoate consumption include thermal decomposition to form radicals of methyl propanoate ( CH2 CH2 (C=O)OCH3 ), butanoyloxy (CH3 CH2 CH (C=O)O ), methyl acetate ( CH2 (C=O)OCH3 ), methoxycarbonyl ( (C=O)OCH3 ), methyl ( CH3 ), ethyl ( C2 H5 ), n-propyl (CH3 CH2 CH2 ), butanoyl (CH3 CH2 CH2 (C=O) ), and methoxy (CH3 O ). Among these species, the butanoyloxy radical is the most favorable because the C-O bond is weaker than the C-C bond. These thermal decomposition reactions become increasingly important and higher temperatures.

H2 C

CO2 + n-C3H7

H3C

H3C

Chapter 5. Results and Discussion

H C H3C H C H3C

O H2 CH2CO + CH3O C C C CH3 O O + X, - XH H C C CO2 + n-C3H7 O O O - CH3 2 H2 H2 C CH2CO + CH3O C H2 C C H2 C CH3 O + X, - XH H C C C CH3 C O O H2 O - CH3 - C2H52 H2 H2 H2C O C C CH3 O O C CH3 + X, - XH H3C H2 C O O - C2H5 H2C O C H2 C CH3 H2 + X, - XH O + Butanoate C CH3 X, - XH H3CMethylC C C CH2 O O H2 CH2O + n-C3H7O C O O H3C H2 C + X, - XH H2 Methyl Butanoate C CH3 C C CH2 CH2O + n-C3H7O C - CH3 O C O H3C H2 - C3H6 - C3H7 + X, - XH O - CH3 H2 O O - C3H6 C C CH3 - C3H7 + X, - XH O H2 + X, - XH O H3C H2 C C C CH3 O C CH3 - CH3 - C2H4 O O C H C CH3 C O H2 H2 O + X, -H2 XH C O H3C C C CH3 C C CH3 - CH3 H C CH3 - C2H4 O C O C O O H2 H H2 +H C C CH3 CH3O + CO C O H H2 + CH2O + C2H3CO or +H O CH32O+ +CH3 CO CO H2 + CH2O + C2H3CO H or C C CH3 O + X, - XH CO2 + CH3 C O H3C H + X, - XH H C C CH3 + X, - XH C O H3C + X, - XH + X, - XH H O HCCO + CH3CHO O H - CH3O + X, - XH or H C C O C C CH3 HCCO + CH3CHO C O CH3CHCHO + CO - CH3O H2C H H O H2C H C or C C H C C CH3 C CH3CHCHO + CO H2C C H O H2C H

Figure 5.20: Primary reaction pathways for methyl butanoate oxidation in the opposed ow diusion ame

89

Chapter 6 Conclusions and Recommendations


6.1 Conclusions

This thesis examined the oxidation of two short chain FAMEs in an opposed ow diusion ame conguration. The primary goal was to compare the emission proles for methyl butanoate and methyl crotonate to determine the role of FAME molecular structure (i.e. unsaturation) in combustion. The experimental results were also used to validate a chemical kinetic model for methyl butanoate, in hopes of better understanding the oxidation of the saturated long chain FAMEs found in biodiesel. Our results indicate that methyl butanoate and methyl crotonate produce similar peak mole fractions for carbon monoxide (CO), carbon dioxide (CO2 ), methane (CH4 ), ethane (C2 H6 ), formaldehyde (CH2 O), and acetaldehyde (CH3 CHO). The methyl butanoate ame produces higher levels of ethylene (C2 H4 ) and propane (C3 H8 ). However, the methyl crotonate ame produces higher concentrations of propene (C3 H6 ), acetylene (C2 H2 ), propyne (C3 H4 ), 1-butene (1-C4 H8 ), 2-butyne (2-C4 H6 ), butane (C4 H10 ), n-pentane (C5 H12 ), and 1,3-butadiene (1,3-C4 H6 ). In addition, the methyl crotonate ame produces methanol (CH3 OH), acrolein (C3 H4 O), acetone (C3 H6 O), and benzene (C6 H6 ) in detectable quantities, while these species were below the detection limit in methyl butanoate ames. The dierences between methyl butanoate and methyl crotonate emissions proles were explained by reaction pathway analysis. Higher concentrations of ethylene (C2 H4 ) and propane (C3 H8 ) in the methyl butanoate ame are attributed to the intermediate n-propyl radical (C3 H7 ), which is derived from the fuel by a series of H abstraction and decomposition reactions. However, the analogous H abstraction and thermal decomposition 90

Chapter 6. Conclusions and Recommendations

91

reactions on methyl crotonate lead to the intermediate 1-propenyl radical (C3 H5 ), which reacts to form propene (C3 H6 ) by H addition and acetylene (C2 H2 ) and propyne (C3 H4 ) by -scission. Acetylene pyrolizes to form higher molecular weight alkenes, alkynes, and aromatic species, thereby rationalizing higher levels of 1-butene (1-C4 H8 ), 2-butyne (2-C4 H6 ), 1,3-butadiene (1,3-C4 H6 ), and benzene in the methyl crotonate ame. The higher levels of methanol (CH3 OH) in methyl crotonate are attributed to reactions involving the intermediate methoxycarbonyl radical (CH3 OCO). This radical is more readily formed in methyl crotonate due to the induction eect of the double bond adjacent to the carbonyl group. Higher levels of acrolein (C3 H4 O) in the methyl crotonate ame are attributed to a series of fuel decomposition and H addition reactions; whereas, the analogous reactions on methyl butanoate lead to the formation of propenal (C3 H6 O). Acetylene, propyne, C4 hydrocarbons, and benzene are known precursors and intermediates in the formation of soot. These compounds were all found in higher levels in the methyl crotonate ame, and it was shown that the double bond in responsible for the formation of these compounds. Therefore, it can be concluded that unsaturated long chain FAMEs will also display a greater tendency to form these soot precursors. Under rich conditions, it can be expected that unsaturated FAMEs will lead to more soot formation than saturated FAMEs. The experimental and modeling results exhibit the benets of the modications performed by members of our Combustion Research Group on the original methyl butanoate oxidation mechanism [8]. The modied mechanism improved upon the original by: i.) including the prediction of C4 hydrocarbons via addition of a C4 sub-mechanism, ii.) improving the prediction of methyl butanoate reactivity via modication of specic reaction rate constants, and iii.) improving the prediction of formaldehyde and carbon monoxide concentration via modication of specic reaction rate constants. The peak concentrations of carbon monoxide (CO), carbon dioxide (CO2 ), methane (CH4 ), ethylene (C2 H4 ), propane (C3 H8 ), and ethane (C2 H6 ) are well predicted by the modied model. However, the model over-predicts the peak concentration of acetylene (C2 H2 ), 1-butene (C4 H8 ), propene C3 H6 , propyne (pC3 H4 ), and formaldehyde (CH2 O), and under-predicts the peak concentrations of 1,3-butadiene (1,3-C4 H6 ) and acetaldehyde (CH3 CHO). The reaction pathways of methyl butanoate oxidation in the opposed ow diusion ame were studied. It is concluded that H abstraction from carbon atoms and unimolecular thermal decomposition are the key consumption reactions, with the former being

Chapter 6. Conclusions and Recommendations

92

more important at low and intermediate temperatures. The primary intermediates are the 2-methyl butanoate and 3-methyl butanoate radicals, which further react to form methyl propenoate and methoxycarbonyl radical, respectively. Unimolecular decomposition of the methoxycarbonyl radical (CH3 CO) leads primarily to CO2 . From a soot suppression standpoint, this is undesirable because two fuel-bound oxygen atoms are bonded to one carbon atom. Ideally, each oxygen atom should bond with a single carbon atom, thereby prevent the carbon atom from being involved in a soot formation reaction. Thus, it can be concluded that the ester moiety will not be highly eective at reducing soot emissions. The initial reaction pathways of methyl crotonate were predicted from the methyl butanoate oxidation mechanism. H abstraction from carbon atoms and unimolecular decomposition are expected to be the major routes of methyl crotonate consumption. In addition, the double bond in methyl crotonate would be susceptible to attack by the biradical O. A validated detailed chemical kinetic model for methyl crotonate will be published in 2006 by our Combustion Research Group.

6.2

Recommendations
the opposed ow diusion ame. However, the model over-predicts the peak concentrations of propene, propyne, acetylene, 1-butene, and formaldehyde. It also under-predicts the concentrations of 1,3-butadiene and acetaldehyde. It is recommended that the model be further rened to provide better agreement for the aforementioned species.

1. The methyl butanoate model well-predicts the concentrations of many species in

2. Currently, a detailed chemical kinetic model for the oxidation of methyl crotonate does not exist. Such a model would provide a better understanding of the oxidation of unsaturated long chain FAMEs. Our Combustion Research Group is in the process of developing this model. 3. Detailed chemical kinetic mechanisms for long chain FAMEs of varying molecular weight and degree of saturation would also be useful. It is recommended that such mechanisms be developed and experimentally validated across a number of welldened combustion platforms. Such a study will be pursued by the author as a PhD student in The Combustion Group.

Chapter 6. Conclusions and Recommendations

93

4. The measurement of higher molecular weight PAH species should be measured in ames of higher molecular weight fuels. The measurement of these species can be accomplished by sequestering the gaseous species on a resin bed, followed by thermal desorption of the gaseous species and analysis using a GC/FID. Such a study will be pursued by the author as a PhD student in The Combustion Group. 5. This study was conducted under non-sooting ame conditions. It is recommended that a study be conducted under sooting conditions to directly quantify the level of soot formation by saturated and unsaturated fuels. Measurements can be obtained using the light extinction technique. Such a study will be pursued by the author as a PhD student in The Combustion Group. 6. It is recommended that the experimental and analytical equipment in The Combustion Research laboratory be upgraded for future studies. The range and accuracy of measured species can be greatly improved by purchasing a new GC/FID integrated with the following: a methanizer for measuring CO2 and CH2 O; a molecular sieve for separating CO and O2 from N2 ; a thermal conductivity detector (TCD) for measuring CO, O2 , and H2 ; and electronically controlled ow controls. 7. The opposed ow diusion ame conditions can be controlled better with the implementation of mass ow meters for measurement of fuel and oxidizer gas ows to the burner. The health and safety of laboratory personnel can be improved by installing an improved ventilation system.

Bibliography
[1] G. G. D. E. Canada, Canadas greenhouse gas inventory 2003 report (Feb. 2006). URL cfm [2] Biodiesel america.org (Feb. 2006). URL http://www.biodieselamerica.org/node/1054 [3] American Society for Testing and Materials, ASTM D 6751 Specication for Biodiesel Fuel Blend Stock (b100) for Middle Distillate fuels, in: ASTM Book of Standards, ASTM, 2003. [4] Ocial site of the National Biodiesel Board (Feb. 2006). URL http://www.biodiesel.org/ [5] T. D. Durbin, J. F. Collins, H. Galdamez, J. M. Norbeck, M. R. Smith, R. D. Wilson, T. Younglove, Evaluation of the eects of biodiesel fuel on emissions from heavy-duty non-road vehicles, Tech. rep., Center for Environmental Research and Technology, College of Engineering, University of California, Riverside (May 2000). [6] T. D. Durbin, K. Cocker, J. F. Collins, J. M. Norbeck, Evaluation of the eects of biodiesel and biodiesel blends on exhaust emission rates and reactivity - 2, Tech. rep., Center for Environmental Research and Technology, College of Engineering, University of California, Riverside (August 2001). [7] U.S.E.P.A. Draft Technical Report EPA420-P-02-001, Tech. rep., United States Environmental Protection Agency (2002). [8] E. Fisher, W. Pitz, H. Curran, C. Westbrook, Detailed chemical kinetic mechanisms for combustion of oxygenated fuels, Symposium (International) on Combustion 28 (2) (2000) 1579 1586. 94 http://www.ec.gc.ca/pdb/ghg/inventory report/2003 report/toc e.

Bibliography

95

[9] S. A. Syed, Oxidation studies of surrogate bio-diesel fuels in opposed ow diusion ames (2005). [10] S. Ga M. Thomson, S. Sarathy, S. Syed, A wide-ranging kinetic modeling study l, of methyl butanoate combustion, sumbitted to Symposium (International) on Combustion (2006) [11] R. McCormick, M. Graboski, T. Alleman, A. Herring, K. Tyson, Impact of biodiesel source material and chemical structure on emissions of criteria pollutants from a heavy-duty engine, Environmental Science and Technology 35 (9) (2001) 1742 1747. [12] U.S. DOE, Biodiesel handling and use guidelines, Tech. rep., U.S. Department of Energy (2004). [13] C. R. Engler, L. A. Johnson, W. A. Lepori, C. M. Yarbrough, Eects of processing and chemical characteristics of plant oils on performance of an indirect-injection diesel engine., JAOCS, Journal of the American Oil Chemists Society 60 (8) (1983) 1592 1596. [14] C. L. Peterson, D. L. Auld, R. A. Korus, Winter rape oil fuel for diesel engines: Recovery and utilization., JAOCS, Journal of the American Oil Chemists Society 60 (8) (1983) 1579 1587. [15] N. J. Schlautman, J. L. Schinstock, M. A. Hanna, Unrened expelled soybean oil performance in a diesel engine., Transactions of the ASAE 29 (1) (1986) 70 73. [16] Olive oil source (Feb. 2006). URL http://www.oliveoilsource.com/images/triglyceride.jpg [17] The hydrogenation of vegetable oils and the production of vegetable ghee, United Nations, New York, 1974. [18] H. Patterson, Hydrogenation of fats and oils, Applied Science, London, 1983. [19] R. Kotrba, Everything under the sun, Biodiesel Magazine (2006) 2934. [20] J. Kinast, Production of biodiesels from multiple feedstocks and properties of biodiesels and biodiesel/diesel blends, Tech. rep., National Renewable Energy Laboratory (2003).

Bibliography

96

[21] F. Ma, M. A. Hanna, Biodiesel production: A review, Bioresource Technology 70 (1) (1999) 1 15. [22] A. Ramadhas, S. Jayaraj, C. Muraleedharan, Use of vegetable oils as i.c. engine fuels - a review, Renewable Energy 29 (5) (2004) 727 742. [23] J. Encinar, J. Gonzalez, J. Rodriguez, A. Tejedor, Biodiesel fuels from vegetable oils: Transesterication of cynara cardunculus l. oils with ethanol, Energy and Fuels 16 (2) (2002) 443 450. [24] S. Jackson, Standard - for a good reason, Biodiesel Magazine (2006) 6062. [25] G. Vicente, M. Martinez, J. Aracil, Integrated biodiesel production: A comparison of dierent homogeneous catalysts systems, Bioresource Technology 92 (3) (2004) 297 305. [26] American Society for Testing and Materials, D 975 Standard Specication for Diesel Fuel Oils, in: ASTM Book of Standards, ASTM, 2003. [27] A. Agarwal, L. Das, Biodiesel development and characterization for use as a fuel in compression ignition engines, Journal of Engineering for Gas Turbines and Power 123 (2) (2001) 440 447. [28] M. Graboski, J. Ross, R. McCormick, Transient emissions from no. 2 diesel and biodiesel blends in a ddc series 60 engine, SAE Special Publications (1179) (1996) 55. [29] M. S. Graboski, R. L. McCormick, Combustion of fat and vegetable oil derived fuels in diesel engines, Progress in Energy and Combustion Science 24 (2) (1998) 125 164. [30] M. Hess, M. Haas, T. Foglia, W. Marmer, Eect of antioxidant addition on nox emissions from biodiesel, Energy and Fuels 19 (4) (2005) 17491754. [31] M. E. Tat, J. H. Van Gerpen, Eect of temperature and pressure on the speed of sound and isentropic bulk modulus of mixtures of biodiesel and diesel fuel, JAOCS, Journal of the American Oil Chemists Society 80 (11) (2003) 1127 1130. [32] A. L. Boehman, D. Morris, J. Szybist, E. Esen, The impact of the bulk modulus of diesel fuels on fuel injection timing, Energy and Fuels 18 (6) (2004) 1877 1882.

Bibliography

97

[33] D. P. Geller, J. W. Goodrum, Eects of specic fatty acid methyl esters on diesel fuel lubricity, Fuel 83 (17-18) (2004) 2351 2356. [34] J. Song, K. Cheenkachorn, J. Wang, J. Perez, A. L. Boehman, P. J. Young, F. J. Waller, Eect of oxygenated fuel on combustion and emissions in a light-duty turbo diesel engine, Energy and Fuels 16 (2) (2002) 294 301. [35] J. M. Simmie, Detailed chemical kinetic models for the combustion of hydrocarbon fuels, Progress in Energy and Combustion Science 29 (6) (2003) 599 634. [36] C. K. Westbrook, F. L. Dryer, Chemical kinetic modeling of hydrocarbon combustion., Progress in Energy and Combustion Science 10 (1) (1984) 1 57. [37] R. Kee, F. Rupley, J. Miller, M. Coltrin, J. Grear, E. Meeks, H. Moat, A. Lutz, G. Dixon-Lewis, M. Smooke, J. Warnatz, G. Evans, R. Larson, R. Mitchell, L. Petzold, W. Reynolds, M. Caracotsios, W. Stewart, P. Glarborg, C. Wang, C. McLellan, O. Adigun, W. Houf, C. Chou, S. Miller, P. Ho, P. Young, D. Young, CHEMKIN Release 4.0.2, Reaction Design, San Diego, CA, 2005. [38] I. Glassman, Combustion, 3rd Edition, Academic Press, San Diego, CA, 1996. [39] N. N. Semenov, Some Problems in Chemical Kinetics and Reactivity, Princeton University Press, Princeton, New Jersey, 1958. [40] S. W. Benson, Kinetics and thermochemistry of chemical oxidation with application to combustion and ames., Progress in Energy and Combustion Science 7 (2) (1981) 125 134. [41] F. Zhao, T. W. Asmus, D. N. Assani, J. E. Dec, J. A. Eng, P. M. Najt, Homogeneous Charge Compression Ignition (HCCI) Engines: Key Research and Development Issues, Society of Automotive Engineers Inc., Warrendale, PA, USA, 2003. [42] P. Dagaut, N. Smoucovit, M. Cathonnet, Methyl acetate oxidation in a jsr: Experimental and detailed kinetic modeling study., Combustion Science and Technology 127 (1997) 275291. [43] P. A. Glaude, W. J. Pitz, M. J. Thomson, R. Pitz, Chemical kinetic modeling of dimethyl carbonate in an opposed-ow diusion ame, Proceedings of the Combustion Institute 30 (1) (2005) 1111 1118.

Bibliography

98

[44] J. P. Szybist, A. L. Boehman, Auto-ignitiion behavior of methyl decanoate, Prepr. Pap.-Am. Chem. Soc., Div. Fuel Chem. 50 (2) (2005) 730. [45] T. C. Bond, K. Sun, Can reducing black carbon emissions counteract global warming?, Environmental Science and Technology 39 (16) (2005) 5921 5926. [46] Anonymous, More data needed to support or disprove global warming theory, Oil and Gas Journal 95 (21) (1997) 75 76. [47] I. M. Kennedy, Models of soot formation and oxidation, Progress in Energy and Combustion Science 23 (2) (1997) 95 132. [48] Catalog of sonics and materials inc. (Feb. 2006) URL www.sonicsandmaterials.com [49] S. M. Schoenung, R. K. Hanson, Co and temperature measurements in a at ame by laser absorption spectroscopy and probe techniques., Combustion Science and Technology 24 (5-6) (1981) 227 237. [50] R. M. Fristrom, Comments on quenching mechanisms in the microprobe sampling of ames, Combustion and Flame 50 (1983) 239242. [51] A. M. Vincitore, S. M. Senkan, Polycyclic aromatic hydrocarbon formation in opposed ow diusion ames of ethane, Combustion and Flame 114 (1-2) (1998) 259 266. [52] A. Sinha, M. J. Thomson, The chemical structures of opposed ow diusion ames of c3 oxygenated hydrocarbons (isopropanol, dimethoxy methane, and dimethyl carbonate) and their mixtures, Combustion and Flame 136 (4) (2004) 548556. [53] S. S. M. Kassem, M. Qun, Chemical structure of fuel-rich 1,2-c2h4cl2/ch4/o2/ar ames: Eects of micro-probe cooling on the sampling of ames of chlorinated hydrocarbons, Combustion Science and Technology 67 (1989) 147157. [54] R. M. Fristrom, Flame structure, McGraw-Hill, New York, 1965. [55] K2BW Environmental Equipment Service Co. (Feb. 2006). URL http://www.k2bw.com/5 c 18.htm

Bibliography [56] B. M. Tissue, Gas chromatography (gc) (Feb. 2006). URL http://elchem.kaist.ac.kr/vt/chem-ed/sep/gc/gc.htm [57] B. M. Tissue, Flame-ionization detectors (d) (Feb. 2006). URL http://elchem.kaist.ac.kr/vt/chem-ed/sep/gc/detector/fid.htm

99

[58] Bartle, The use of helium ionization detector for gas chromatographic monitoring of trace atmospheric components, Journal of High Resolution Chromatography 21 (1998) 7580. [59] S. Tsang, L. Graham, L. Shen, X. Zhou, T. Lanni, Simultaneous determination of carbonyls and no2 exhausts of heavy-duty diesel trucks and transit buses by hlpc following 2,4-dinitrophenylhydrazine catridge collection, Environmental Science and Technology 38 (22) (2004) 59685976. [60] E. Grosjean, P. G. Green, D. Grosjean, Liquid chromatography analysis of carbonyl (2,4-dinitrophenyl)hydrazones with detection by diode array ultraviolet spectroscopy and by atmospheric pressure negative chemical ionization mass spectrometry, Analytical Chemistry 71 (9) (1999) 18511861. [61] American Society for Testing and Materials, D5197-03 Standard Test Method for Determination of Formaldehyde and Other Carbonyl Compounds in Air (Active Sampler Methodology), in: ASTM Book of Standards, Vol. 11, ASTM, 2003. [62] B. M. Tissue, High-performance liquid chromatography (hplc)/hplc.htm (Feb. 2006). URL http://elchem.kaist.ac.kr/vt/chem-ed/sep/lc/hplc.htm [63] D. Bradley, A. Entwistle, Deterimination of the emissivity, for total radiation, of small diameter platinum-10% rhodium wires in the temperature range 600-450 degrees C, British Journal of Applied Physics 12 (1961) 708711. [64] R. Design, CHEMKIN Theory Manual, Reaction Design, 4th Edition (April 2005). [65] C. K. Westbrook, W. J. Pitz, LLNL combustion chemistry group - mechanisms (Feb. 2006). URL http://www-cms.llnl.gov/combustion/combustion2.html

Bibliography

100

[66] A. Demirbas, Direct route to the calculation of heating values of liquid fuels by using their density and viscosity measurements, Energy Conversion and Management 41 (15) (2000) 1609 1614. [67] National Institute of Standards and Technology Scientic and Technical Databases - Chemical Kinetics (Feb. 2006). URL http://www.nist.gov/srd/chemkin.htm [68] W. Tsang, R. Hampson, Chemical kinetic databse for combustion chemistry: Part I. Methane and related compounds, Journal of Physical Chemical Reference Data 15 (1986). [69] S. G. Davis, C. K. Law, H. Wang, Propyne pyrolysis in a ow reactor: An experimental, rrkm, and detailed kinetic modeling study, Journal of Physical Chemistry A 103 (30) (1999) 5889 5899. [70] Y. Hidaka, T. Nakamura, H. Tanaka, A. Jinno, H. Kawano, T. Higashihara, Shock tube and modeling study of propene pyrolysis, International Journal of Chemical Kinetics 24 (9) (1992) 761 780.

Appendix A Flow Rates of Fuel and Oxidizer

101

Appendix A. Flow Rates of Fuel and Oxidizer

102

Flow rates of fuel and oxidizer


Methyl butanaote (mb) Fuel or oxidizer stream Nitrogen (gas) Methyl butanaote (liq.) Air (gas) Oxygen (gas) Fuel Fuel Oxidizer Oxidizer Rotameter reading 65 SS NA 41 SS 65 SS

Flow rate 3.11 slpm 0.72 ml/min 2.24 0.826

Density of mb = 0.8984 g/ml; Molecular weight of mb = 102.13 g/mol Therefore, molar flow rate of mb = 6.32 E-3 mol/min Molar flow rate of N2 in fuel stream = 0.127747 mol/min Mole fraction of N2 in fuel stream = 95.28 % Mole fraction of mb in fuel stream = 4.72% Similarly, for the oxidizer stream the mole fractions of nitrogen and oxygen are obtained. Mole fraction of N2 in oxidizer stream = 57.75% Mole fraction of O2 in oxidizer stream = 42.25%

Appendix A. Flow Rates of Fuel and Oxidizer

103

Methyl crotonate (mc) Fuel or oxidizer stream Nitrogen (gas) Methyl crotonate (liq.) Air (gas) Oxygen (gas) Fuel Fuel Oxidizer Oxidizer Rotameter reading 65 SS NA 41 SS 65 SS

Flow rate 3.11 slpm 0.67 ml/min 2.24 0.826

Density of mc = 0.945 g/ml; Molecular weight of mc = 100.12 g/mol Therefore, molar flow rate of mc = 6.32 E-3 mol/min, same as that of mb Molar flow rate of N2 in fuel stream = 0.127747 mol/min Mole fraction of N2 in fuel stream = 95.28 % Mole fraction of mc in fuel stream = 4.72% Similarly, for the oxidizer stream the mole fractions of nitrogen and oxygen are obtained. Mole fraction of N2 in oxidizer stream = 57.75% Mole fraction of O2 in oxidizer stream = 42.25%

Appendix B Structures of Chemical Compounds

104

Appendix B. Structures of Chemical Compounds

105

OH

propene

acetylene
O

methanol
O

methyl crotonate
O

acetaldehyde
O

allene

carbon monoxide

CH3

methyl radical

carbon dioxide methyl butyrate

acetone

CH2:

methylene radical

formaldehyde acrolein benzene


CH4

1,3-butadiene

methane

propene

2-butyne

CH2

n-propyl radical

ethane

propyne

butane

H C

iso-propyl radical

ethylene

propane

n-pentane

O CH C

propenyl radical

acetylene

1-butene

methoxy carbonyl

Table B.1: Structures of relevant chemical species

Appendix C Sample Chromatograms

106

Appendix C. Sample Chromatograms

107

Figure C.1: Example of HPLC chromatogram of DNPH derivatives

Appendix C. Sample Chromatograms

108

Figure C.2: Example of GC chromatogram for the HP Plot Column

Appendix C. Sample Chromatograms

109

1,3-Butadiene
Column:
P/N: Carrier: Oven:

DB-624

25 m x 0.20 mm I.D., 1.12 m 128-1324 Helium at 1.0 mL/min -20C for 3 min -20 - 20C at 4/min 20 - 200C at 8/min 200C for 10 min Injector: Split 1:150, 250C 0.5 L injection Detector: FID, 250C Agilent Technologies wishes to thank DCG Industries (Pearland, TX) for providing this chromatogram. 2, 3 7

11 8, 9 10 12 13

20

Refined Butadiene Standard Component 1. Acetylene 2. Propane 3. Propylene 4. Propadiene (allene) 5. Propyne (methylacetylene) 6. Cyclopropane 7. Isobutane 8. Butene-1 9. Isobutylene 10. n-Butane 11. 1,3-Butadiene 12. trans-2-Butene 13. cis-2-Butene 14. 1-Butyne (ethylacetylene) 15. 1,2-Butadiene 16. 3-Methyl-1-butene 17. Isopentane 18. Pentene-1 19. n-Pentane 20. 2-Butyne (dimethylacetylene) 21. trans-2-Pentene 22. Isoprene 23. cis-2-Pentene 24. trans-1,3-Pentadiene 25. cis-1,3-Pentadiene 26. Benzene 27. Toluene 28. Dimer (4-vinylcyclohexene-1)

Gravimetric concentration (PPM) 20.7 19.8 296 21.1 21.0 20.0 506 999 495 494 balance 28 442 1946 20.2 28.9 19.8 50.1 29.8 50.1 150 5.57 20.0 13.9 13.8 7.73 20.3 20.2

5 4

6 14 15 16

17

19 18

22 23 21

26 24 25

27

C777

10 Time (min)

15

20

Figure C.3: Example of GC chromatogram for the DB 624 Column

Appendix D Sample Calculations

110

You might also like