150
University of Groningen Experimental study of the combustion properties of methane/hydrogen mixtures Gersen, Sander IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below. Document Version Publisher's PDF, also known as Version of record Publication date: 2007 Link to publication in University of Groningen/UMCG research database Citation for published version (APA): Gersen, S. (2007). Experimental study of the combustion properties of methane/hydrogen mixtures. s.n. Copyright Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons). Take-down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum. Download date: 26-03-2020

University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

University of Groningen

Experimental study of the combustion properties of methane/hydrogen mixturesGersen, Sander

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite fromit. Please check the document version below.

Document VersionPublisher's PDF, also known as Version of record

Publication date:2007

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):Gersen, S. (2007). Experimental study of the combustion properties of methane/hydrogen mixtures. s.n.

CopyrightOther than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of theauthor(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons thenumber of authors shown on this cover page is limited to 10 maximum.

Download date: 26-03-2020

Page 2: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

RIJKSUNIVERSITEIT GRONINGEN

Experimental study of the combustion properties of methane/hydrogen mixtures

Proefschrift

ter verkrijging van het doctoraat in de Wiskunde en Natuurwetenschappen aan de Rijksuniversiteit Groningen

op gezag van de Rector Magnificus, dr. F. Zwarts, in het openbaar te verdedigen op

vrijdag 7 december 2007 om 14:45 uur

door

Sander Gersen geboren op 2 oktober 1976

te Gouda

Page 3: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Promotor: Prof. dr. H.B. Levinsky Copromotor: Dr. A.V. Mokhov Beoordelingscommissie: Prof. dr. ir. R. Baert Prof. dr. H.C. Moll Prof. dr. ir. Th.H. van der Meer ISBN 978-90-367-3254-3 ISBN 978-90-367-3255-0 (electronic version)

Page 4: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Experimental study of the combustion properties of methane/hydrogen

mixtures

Sander Gersen

Page 5: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

The work described in this thesis was performed in the Laboratory for Fuel and

Combustion Science at the University of Groningen, Nijenborgh 4, 9747 AG

Groningen, the Netherlands. This project is supported with a grant of the Dutch

Program EET (Economy, Ecology, Technology) a joint initiative of the Ministries of

Economic Affairs, Education, Culture and Sciences and of Housing, Spatial Planning

and the Environment. The program is run by the EET Program Office, SenterNovem.

S.Gersen,

Experimental study of the combustion properties of methane/hydrogen mixtures,

Proefschrift Rijksuniversiteit Groningen (2007)

Page 6: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

I

Table of contents Introduction 5

Chapter 1 : Combustion properties of homogeneous reacting gas mixtures

1.1 Motivation to study the combustion properties of CH4/H2 gas mixtures 8

1.2 Governing equations for a homogeneous reacting gas mixture in a closed

gas system 16

1.3 Laminar premixed flames 18

1.3.1 Governing equations for a one-dimensional laminar flame 21

1.4 Chemical mechanisms 22

Chapter 2: The Rapid Compression Machine, Experimental Techniques, Procedures and Setup

2.1 Background 27

2.2 Design and Operations 38

2.2.1 Experimental System 32

2.2.2 Gas filling system and filling procedure 33

2.2.3 Instrumentation and data acquisition 34

2.3.4 Determination autoignition delay time 35

2.3.5. Temperature determination 36

Appendix A.1 40

Appendix A.2 41

Appendix A.3 42

Appendix A.4 43

Appendix A.5 44

Chapter 3: High-pressure autoignition delay time measurements in methane/hydrogen fuel mixtures in a Rapid Compression Machine

3.1 Introduction 49

3.2 Experimental approach 51

3.3 Numerical simulation and analysis of experimental data 52

Page 7: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

II

3.3.1 Chemical mechanisms 52

3.3.2 Numerical simulations 53

3.4 Results and discussion 56

3.5 Comparison of experimental results with numerical simulations 63

3.6 Summary and conclusions 68

Chapter 4: One-dimensional laminar flames, Experimental Techniques,

Procedures and Burner Setup

4.1 General introduction 73

4.2 Burner 76

4.3 Gas handling system 77

4.4 Extractive probe sampling system 79

4.5 Estimate of the conversion of C2H2 and HCN during sampling 80

4.6 Laser absorption spectroscopy 82

4.6.1 Theory 82

4.6.2 Wavelength Modulation Absorption Spectroscopy (WMAS) 83

4.7 Experimental setup for Tunable Diode Laser Absorption Spectroscopy 87

4.7.1 Experimental procedure TDAS measurements of acetylene 88

4.8 Experimental procedure WMAS with second harmonic detection 90

4.8.1 HCN measurements 90

Chapter 5: Extractive Probe Measurements of acetylene in atmospheric pressure

fuel-rich premixed methane/air flames

5.1 Introduction 100

5.2 Experimental 101

5.3 Results and discussion 102

5.4 Conclusions 107

Page 8: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

III

Chapter 6: HCN formation and destruction in atmospheric pressure fuel-rich

premixed methane/air flames

6.1 Introduction 110

6.2 Experimental 112

6.3 Results and discussion 113

6.4 Conclusions 120

Chapter 7: The effect of hydrogen addition to rich stabilized methane/air flames

7.1 Introduction 124

7.2 Experimental 126

7.3 Results and discussion 127

7.3.1 HCN profiles 128

7.3.2 C2H2 profiles 131

7.4 Conclusion 132

Summary 136

Samenvatting 140

Dankwoord 144

Page 9: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

IV

Page 10: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Inroduction

5

Introduction

Combustion is mankind’s oldest technology. Nowadays the combustion of fossil

fuels provides more than 80% of the world’s energy, and is used for electric power

generation, domestic heating, transportation and many other processes. A negative

aspect of fossil fuels is that during combustion not only heat is generated, but also

pollutants such as soot and NOx. Moreover, the combustion of fossil fuels disturbs the

atmospheric CO2 balance, which is believed to contribute to global warming.

Stringent emission regulations and the expectation that the known fossil fuel

reserves will be exhausted within this century, forces combustion researchers to find

methods to reduce pollutant emission, improve the efficiency of combustion

equipment and to utilize renewable energy sources, such as biogas and hydrogen, as

alternative fuels. However, the currently available renewable energy sources are

insufficient to satisfy the world’s energy consumption. In a sustainable economy,

hydrogen, either from electrolysis of water from sustainable generated electricity

(wind, water) or from biomass, can fulfill a role as energy carrier. Yet at present, there

is no sustainable hydrogen production, nor is there widespread energy conversion

technology to utilize hydrogen as a fuel. To avoid the necessity of large investment in

new hydrogen utilization equipment, the addition of hydrogen to natural gas could be

a first step towards the wide-scale introduction of hydrogen into the energy

infrastructure. However, since the combustion properties of hydrogen differ in many

respects from those of natural gas, the allowable fraction of hydrogen in natural gas

may be limited by the deteriorating performance of gas combustion equipment such as

spark-ignited engines, burners and turbines to hydrogen-enriched natural gas. For

example, increased knock in gas engines, causing extensive damage to the machines,

or unacceptable increases in NOx formation from combustion equipment, both caused

by the presence of hydrogen in the fuel, are clearly unwanted side-effects and must be

avoided.

To investigate these practical consequences of the changes if fuel composition

effectively, it is necessary to study the changes in the underlying physical and

chemical processes that are responsible for the combustion behavior of natural gas

when hydrogen is added. Gaining fundamental insight into these consequences using

practical combustion devices is difficult, since the experimental conditions are

generally poorly defined, complicating the interpretation of the data. For this reason,

Page 11: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Introduction

6

devices as shock tubes, rapid compression machines (RCM) and one-dimensional

flame burners have been developed to enable the study of combustion under well-

defined conditions. The insights gained from such studies permit the analysis of the

behavior of broad groups of practical combustion equipment, and are also

indispensable for the design of new combustion equipment. Furthermore, data from

these well-defined studies aid the development of methods for modeling complex

combustion phenomena.

The objective of this thesis is to investigate potential changes in the combustion

properties of methane caused by the addition of hydrogen to the fuel. Specifically, the

ignition properties of methane/oxygen and methane/hydrogen/oxygen mixtures are

studied by measuring auto ignition delay times in a rapid compression machine

(RCM) at conditions relevant to knock in gas engines (950<T<1100K and

10<P<70bar). In addition, insight into changes in soot and NOx formation in methane

flames is gained by measuring the spatial profiles of C2H2 and HCN in atmospheric

pressure, one-dimensional CH4/air and CH4/H2/air flames. All measurements are

compared with the results of numerical calculations designed to predict the behavior

of these experimental systems.

Page 12: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

7

CHAPTER 1

Combustion properties of homogeneous reacting gas mixtures

Page 13: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

8

1.1. Motivation to study the combustion properties of CH4/H2 gas mixtures

Consider a homogeneous (premixed) fuel/oxidizer gas mixture. A premixed

mixture is characterized by the equivalence ratio, ϕ, which expresses the ratio of fuel

and oxidizer the unburned mixture. This is given by,

stfOxidizerFuel 1.

][][

=ϕ , (1.1)

where the amounts [fuel] and [oxidizer] can be expressed in molar, volume or mass

units, and fst is the ratio of fuel to oxidizer under stoichiometric conditions using the

same units. Here we will generally use moles or mole fraction as units. A mixture is

said to be stoichiometric (ϕ=1) when fuel and oxidizer are present in the ratios

prescribed by the balanced chemical reaction for combustion:

OHnxCOOnxHC nx 222 2)

4( +→++ , (R1.1)

If the oxidizer in the unburned mixture is in excess, the mixture is said to be fuel-lean

(ϕ<1), while the mixture is called fuel-rich (ϕ>1) when an excess of fuel in the

unburned mixture is present. The gas mixture can remain unreacted, such as in fuel-air

mixtures at room temperature in the absence of an ignition source, or the fuel and

oxidizer react (combust) to form products. Combustion can take place either in a

flame (a reaction front propagates subsonically through the mixture) or in a non-flame

mode (“homogeneous combustion”, reaction occurs simultaneously everywhere in the

mixture). To understand which mode of combustion takes place under a given set of

conditions (temperature, pressure, equivalence ratio), it is necessary to study the

chemical processes in the system in detail. The overall combustion process can be

described by reaction (R1.1). However, it is unrealistic to think that combustion

proceeds via this single reaction because it would require breaking high-energy bonds,

which at room temperature makes this reaction extremely slow. Instead, combustion

occurs in a sequential process involving many reactive intermediate species.

To illustrate this process, we first consider a H2-O2 gas mixture. The conversion of

hydrogen and oxygen to water starts with the formation of reactive species (radicals)

Page 14: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

9

to initiate a chain of reactions [1]. The reactions in which radicals are formed from

stable species are called chain-initiation reactions, and an example of a chain-

initiation reaction is the (endothermic) dissociation reaction:

molekJMHMH /43622 −+=+ , (R1.2)

The H radicals formed in reaction (R1.2) can react further with oxygen molecules,

forming two new radicals, OH and O,

molekJOOHOH /6.702 −+=+ . (R1.3)

This reaction (R1.3), in which two radicals are created for each radical consumed is

called a chain-branching reaction and is crucially important in combustion processes.

The formation of the radicals OH and O can lead to further chain branching via

molekJHOHHO /2.82 −+=+ . (R1.4)

In addition, there are reactions in which the number of radicals does not change, such

as

molekJHOHHOH /21.6322 ++=+ , (R1.5)

which are called chain-propagating reactions. The reactions in which radicals react to

stable species without forming another radical are called chain-terminating reactions,

as in

molekJMOHMOHH /2.4992 ++=++ . (R1.6)

Although not strictly chain terminating, since HO2 is a radical, the reaction

molekJMHOMOH /25.20822 ++=++ (R1.7)

Page 15: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

10

is often considered chain terminating because, compared with the “flame radicals”, H,

O and OH, the HO2 radical is relatively unreactive. At low pressures, reaction (R1.7)

is an important chain-terminating reaction because the mildly reactive HO2 radicals

diffuse to the wall, where they react at the surface. Summing the chain-branching and

chain-propagating reactions (R1.3+R1.4+R1.5+R1.5) results in

molekJOHHOHH /62.47233 222 ++→++ , (R1.8)

from which we can see that starting with one radical, three radicals are formed from

the reactants in this simplified mechanism.

For quantitative description of chemical processes the rate of change of the

species concentrations (formation and consumption) should be determined. For

species A in an arbitrary bimolecular elementary reaction,

,dDcCbBaAf

r

k

k+⇔+ (R1.9)

this is expressed as:

,][][][][ dcr

baf DCkBAk

dtdA

+−= (1.2)

where A,B,C,D denote the different species in the reaction, a,b,c,d are the

stoichiometric coefficients of species A,B,C,D, respectively, and kf and kr represent

the forward and reverse rate coefficient of the reaction. For example, the rate of

change of the oxygen radical in reaction (R1.3) is expressed as,

]][[]][[][ 3.12

3.1 OOHkOHkdtOd R

rRf +−= . (1.3)

The rate coefficients kf and kr are connected through the equilibrium constant Kw. The

reaction rate constant k of a reaction is generally assumed to have a modified-

Arrhenius temperature dependence,

expb AEk ATRT

⎛ ⎞= −⎜ ⎟⎝ ⎠

, (1.4)

Page 16: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

11

where A is the pre-exponential factor, b the temperature exponent, T the temperature,

R the universal gas constant and Ea the activation energy, which corresponds to the

energy barrier that has to be overcome during reaction. Here should be mentioned that

the activation energy is always higher than the heat of the formation. Generally, the

more exothermic a reaction is, the smaller the activation energy.

As can be seen from reactions (R1.3-R1.5), the rate of formation of the

important free radicals )( OOHH ++ is proportional to the concentration [n] of the

radicals with some coefficient α. The rate of consumption of radicals (chain

termination) is proportional to the concentration [n] as well (R1.7), with some rate β;

the rate of chain initiation, denoted as γ, is independent of the concentration [n]

(R1.2). Thus, in generalized form, the rate of change of the concentration of free

radicals can be expressed as,

])[(][ ndtnd βαγ −+= , (1.5)

From equation (1.5) three different scenarios can be derived, which are presented

schematically in figure 1.1 a [2].

Figure 1.1a) Schematic of the growth of the concentration of free radicals [n] in time. b) Schematic of the growth of free radicals [n] in time for the cases with and without heat release.

Page 17: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

12

For the condition α>β the concentration [n] increases exponentially in time, and

ignition takes place. When α<β, d[n]/dt becomes zero, and no exponential growth of

free radicals occurs (no ignition). The condition α=β results in a linear growth of the

concentration of free radicals, and defines the ignition limit. Equation (1.5) shows that

the growth of the concentration of free radicals is determined by the competition

between the chain branching (R1.3) and chain terminating reactions (R1.7). A very

important parameter in this competition is the temperature, since the chain branching

reaction (R1.3) has large activation energy Ea [3] while that of the chain terminating

reaction (R1.7) is small [4]. Thus, the value of α (chain branching) is strongly

dependent upon the temperature, and β (chain terminating) is more or less

independent of the temperature. At low temperatures the endothermic chain branching

reaction will not proceed rapidly, so the value of α is much smaller than β (α<β), and

ignition does not occur. Increasing the temperature results in an increase in α, while β

remains unchanged; at sufficiently high temperatures α will be larger than β and

ignition occurs. The temperature during the early period of the ignition process

remains more or less constant because the heat release from the branching and

propagating reactions (R1.8) is small, but as the radical concentration grows,

exothermic reactions such as (R1.6) and (R1.7) will produce substantial quantities of

heat. If the heat produced by the exothermic reactions in system exceeds the rate of

heat loss to the surrounding, the temperature in the system will rise. Since the rate of

reaction, and thus the rate of heat release, grows exponentially with temperature, the

overall reaction will auto-accelerate, that is, the system will “explode”. The time

before explosion takes place is called autoignition delay time (figure 1.1b). In this

example, if no heat accumulates in the system (T=constant), no auto-acceleration of

the reaction rate takes place; in this case we speak of “non-explosive” reactions, both

situations are shown in figure 1.1b.

As described above, the dominance of chain branching reaction (R1.3)

characterizes the high temperature regime, while at low and intermediate temperatures

the in essential chain terminating reaction (R1.7) competes effectively with reaction

(R1.3) [5]. Since the rate of a three-body reaction increases with pressure much faster

than the rate of a two-body reaction, there exists a pressure above which reaction

(R1.7) exceeds the rate of the competing reaction (R1.3). Reaction (R1.7) is only a

chain terminating reaction when the produced HO2 radicals will diffuse to the wall

Page 18: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

13

and recombine to stable molecules without having undergone a reaction. At high

pressures, species collide much more frequently. Therefore, at sufficiently high

pressures the HO2 radicals will be frequently interrupted in its path to the walls by

reacting with H2 molecules to produce H and H2O2 (R1.10) [5],

molekJHOHHHO /45.722222 −+=+ . (R1.10)

The H radicals so produced can contribute to chain branching via (R1.3) or will

generate more HO2 via reaction (R1.7) and H2O2 itself can contribute to branching via

[5],

molekJMOHMOH /6.214222 −+=+ . (R1.11)

Thus at high pressure and moderate temperatures reaction (R1.7) will dominate over

(R1.3), and the number of active centers will grow (α>β) via the sequence of

reactions (R1.7), (R1.10), and (R1.11).

Instead of heating the “cold” gas mixture by the heat release of exothermic

reactions as in this example of a closed homogeneous system, in flames, the mixture

is heated by conduction from the hot flame gases. Furthermore, radicals needed to

decompose the fuel are also transported from the high temperature region of the

flame. As in the closed system described above, the rate of formation of radicals

controls the overall rate of reaction in flames. For flames, however, the chain

initiation and chain-terminating reactions are less important in the

formation/destruction of radicals (α>>β), and the reactions (R1.3)-(R1.5), responsible

for the growth of the radical pool, dominate the overall reaction rate in H2-O2 flames

[6].

The combustion chemistry of hydrocarbon fuels is much more complicated than

that of hydrogen. As an example, we consider a CH4-O2 mixture. Under flame

conditions (α>>β), the most important chain branching reaction in the oxidation of

methane is also reaction (R1.3) [6]. After radicals are transported into the unburned

gas mixture, methane is attacked by the radicals, as in

molekJHCHHCH /13234 −+→+ . (R1.12)

Page 19: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

14

As can be seen from figure 1.2, the rate coefficient of reaction (R1.12) [7] is much

larger than that for reaction (R1.3) [3]. Thus reaction (R1.12) competes effectively

with reaction (R1.3) for H atoms and reduces the chain branching rate.

Figure 1.2. Reaction rate expressions for reaction R1.3 [3] and reaction R1.12 [7].

Furthermore, the very reactive H radical is replaced in reaction (R1.12) by the

unreactive CH3 radical. These two processes result in a slow conversion of methane

and contributes to the low burning velocity of methane (40 cm/s) as compared to

hydrogen (340 cm/s) [8].

If the CH4-O2 mixture under consideration is at moderate or low temperature (T

below roughly 1100K), chain branching reaction (R1.3) is too slow to provide a

sufficient branching rate for autoignition, and a different reaction path dominates [9].

These paths are extremely complex and strongly dependent upon temperature and

pressure [10]. At sufficiently high pressures, reactions involving the radical HO2

become important in the low temperature regime, for example [9],

molekJOHCHHOCH /5.8522324 −+→+ . (R1.13)

The oxidation of the CH3 formed, and the development of the radical pool, is

complicated and slow. This process dominates most of the ignition delay period and is

Page 20: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

15

characterized by the accumulation of significant amounts intermediate species such as

H2, C2H6, CH2O and H2O2. The decomposition of H2O2 via reaction (R1.11) and the

oxidation of CH2O via a sequence of reactions lead to a sharp increase in the

concentration of free radicals [10] and ultimately ignition occurs (α>β) [10].

The H (R1.12) and HO2 (R1.13) scavenging reactions compete effectively with

R1.3 and R1.10 respectively, and thus effectively reduce the chain branching rate in

the CH4-O2 system. This, together with the formation of the relatively unreactive CH3

radical, contributes to the fact that CH4-O2 mixtures tend to auto ignite slower than

H2-O2 mixtures [6], illustrated in figure 1.3.

Figure 1.3. Computed autoignition delay times for stoichiometric H2/air and CH4/air

mixtures at P=30 bar. Calculations were made using the GRI-Mech 3.0 mechanism [11].

Besides their effects on the combustion properties, such as burning velocity and

ignition, the differences in the combustion chemistry of methane and hydrogen have

significant consequences for pollutant formation. One of the main consequences is

that during the combustion of methane (and other hydrocarbons) carbon containing

pollutant species like soot, HCN and CO are formed, while the only pollutant from

hydrogen combustion is NOx. Moreover, the formation of NO in methane combustion

is different than that from H2 combustion; in methane flames an additional mechanism

exist that produces NO via the hydrocarbon intermediate CH [12]:

Page 21: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

16

NOproductsNCHHCNNCN

⇔⇔+?)?,(

2 . (R1.14)

This mechanism is particularly important under fuel rich conditions.

A challenging task is to understand the possible changes in the combustion

chemistry caused by addition of hydrogen to methane and how this affects

combustion properties like pollutant formation and ignition delay. Since there is a

clear distinction between the chemistry in flames and ignition, it is necessary to study

both. To gain understanding in the underlying chemical kinetics and physical

processes involved in these two kinds of combustion process, it is necessary to

analyze the combustion processes quantitatively by solving the governing equations

with detailed chemical mechanisms.

1.2 Governing equations for a homogeneous reacting gas mixture in a closed

system

The time-dependent behavior of a closed system containing a reacting gas

mixture is described by the system of the conservation equations for mass and energy.

Since no mass can be formed or destroyed by chemical reaction, the total mass of the

closed systems remains constant over time:

( ) 0,1

Kd V d VYkdt dt k

ρ ρ= ==∑ (1.6)

where ρ is the overall mass density, V is the system volume and Yk is the mass fraction

of k-th component in the gas mixture. The mass fractions of individual species change

in time according to

dYk Wk kdtρ ω= , k = 1…….K (1.7)

Page 22: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

17

where ωk and Wk are the molar chemical production rate per unit of volume and

molecular weight of the k-th species, respectively. The density is related to

temperature T, pressure p and composition through the ideal gas equation of state:

p W

RTρ = , (1.8)

where 1

1/K

k

k k

YWM=

⎛ ⎞= ⎜ ⎟

⎝ ⎠∑ is the average molecular weight of the mixture. The system

(1.6) – (1.8) consists of (K+1) linearly independent equations and contains (K+3)

unknown parameters: ρ, p, T, V and Yk. Since the system contains more unknowns

than equations, it can be solved only when two unknown parameters (for example, the

measured temperature and pressure) are used as input. The number of input

parameters can be decreased to one if the energy conservation equation is added to the

system. The energy conservation equation can be derived from the first law of

thermodynamics, which states that heat δQ added to the system is equal to the sum of

the change of its internal energy dU and the work PdV of the system against an

external force:

dU PdV Qδ+ = . (1.9)

Equation (1.9) can be rewritten as

lossQdtdQ

dtdVP

dtdU •

==+ . (1.10)

The internal energy of the mixture is given by

1

KU VY uk k

kρ=

=∑ , (1.11)

where uk is a specific energy of the k-th component. After differentiating expression

(1.11) and substituting in (1.10) one receives

Page 23: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

18

1.1

1

d KdTC p u W qv k k k lossdt dt k

ρω

ρ

⎛ ⎞⎜ ⎟⎝ ⎠+ + =

=∑ , (1.12)

where qloss = lossQ•

/(ρV) is the heat loss per unit of mass and Cv is the specific heat of

system at constant volume.

For an adiabatic mixture of inert gases (ωk = 0 and qloss = 0), equation (1.12) can be

easily integrated, resulting in the following expression

∫ ⎟⎟⎠

⎞⎜⎜⎝

⎛=

T

T

vTdT

RWC

00

lnρρ . (1.13)

It is common to use the ratio of molar heat capacities at constant pressure and constant

volume, γ, and a specific volume v, instead of Cv and ρ. In this case, equation (1.13)

can be rewritten as

∫ ⎟⎠

⎞⎜⎝

⎛=−

T

TVV

TdT

0

0ln)1(

. (1.14)

Several simulation programs have been developed to solve the set of

governing equations. The program used in this study is SENKIN [13], and runs in

conjunction with pre-processors from the CHEMKIN library [14], which incorporate

the chemical mechanism and thermodynamic properties.

1.3 Laminar premixed flames

Flat laminar premixed flames are ideally suited for combustion research, since

the one-dimensional character offers great advantages for modeling and unambiguous

model-experiment comparison. Moreover, the structure of these flames is

representative for many practical flames. The structure of laminar premixed flames

Page 24: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

19

can be divided in three zones (figure 1.4), the preheat zone, the flame front (reaction

zone) and the post-flame zone (burned-gas zone). In the preheat zone the unburned

gas mixture is heated by conduction and diffusion of species from the flame front; this

zone can be considered as chemically inert. The flame front, located downstream of

the preheat zone is a thin zone (in the order of 1 mm at atmospheric pressure) in

which the fuel is rapidly oxidized by radicals from the post-flame zone as described

above, leading to a steep gradients in temperature and species concentrations. The

flame front is rich in radicals and intermediate species. Although the temperature and

major species in the post-flame zone are close to their equilibrium value, the

concentrations of minor species can differ substantially from their equilibrium value.

In the post-flame zone, the system goes to equilibrium predominantly via radical-

recombination reactions such as (R1.6).

Figure 1.4. Schematic illustration of the structure of a premix one-dimensional flame.

Premixed flat flames can be characterized by the “free-flame” laminar burning

velocity, vL. In the laboratory system, where the cold gas moves with velocity vu, the

flame front propagates with velocity vu-vL. We can consider three situations regarding

the stability of a idealized one-dimensional flame. If the cold gas velocity is larger

than the laminar burning velocity, vu>vL, the flame front propagates upstream. When

Page 25: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

20

the burning velocity, vL is equal to the velocity of the unburned gasses, (v=vL) the

flame front is stationary in space, and if vu<vL the flame front will be convected

downstream. The laminar burning velocity and the temperature of the burned gas are

completely determined by the properties of the unburned mixture, such as the

equivalence ratio, temperature and the identity of the fuel [2]. Figure 1.5 shows the

interaction of the idealized 1-D flame with a porous-plug burner. Figure 1.5a

illustrates a flat flame stabilized on a burner where the unburned gas velocity is set

equal to the free-flame laminar burning velocity (vu=vL). In this situation, all heat

generated during combustion is transferred completely into the gas mixture and the

flame is essentially adiabatic (neglecting flame radiation). Lowering the unburned gas

velocity vu causes propagation of the flame front towards the burner surface. Since the

porous plug is too dense to allow propagation of the flame into the burner, the flame is

stopped in its upstream propagation. In this case, the flame transfers heat to the burner

by conduction, lowering the flame temperature and thus lowering the actual burning

velocity of the flame vL'. The flame transfers enough heat to the burner to reach a

stationary situation (vu= vL'), illustrated in figure 1.5b. This type of flame is called a

burner-stabilized flame.

Figure 1.5. a) Adiabatic flat flame (freely propagating flame) b) Burner-stabilized flat flame

Further decrease in the unburned gas velocity results in increasing heat loss and drop

in temperature; ultimately the temperature drops to such a level that α<β and the

flame extinguishes.

Page 26: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

21

1.3.1. Governing equations for a one-dimensional laminar flame

The description of one-dimensional laminar flames is based on the conservation

equation for mass, species mass fraction and energy. Using the assumptions that one-

dimensional laminar flames are: (1) stationary (all flame parameters are independent

of time), (2) the system is at constant pressure and (3) effects due to viscosity,

radiation and external forces are negligible [2,15], the conservation equations

governing the behavior of these flames can be summarized as follows:

overall conservation of mass

0)(==

dxdM

dxvd ρ , (1.15)

where v is the mass averaged flow velocity, x is the distance along the line normal to

the burner surface and M is called the mass flux.

conservation of species

( ( ))d Y V vk k Wk kdxρ

ω+

= , k=1….K (1.16)

where Yk is the mass fraction and Vk is the diffusion velocity, which accounts for the

effect of molecular transport due to concentration gradients of the kth species [2,16].

Since mass can neither be destroyed nor formed in chemical reactions it follows from

(1.15) and (1.16) that,

( ( )) ( ) 01 1

K Kd Y V v d vk k Wk kdx dxk k

ρ ρω+

= = == =∑ ∑ k=1…K. (1.17)

Addition of the ideal gas equation of state (1.8) to the system of equations (1.15, 1.16)

results in a system containing (K+1) linear independent equations. Assuming that the

diffusion velocity Vk is a known function of temperature and species concentrations,

Page 27: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

22

the system contains (K+2) unknown parameters (T, ρ, v and Yk). Therefore an

additional equation should be introduced to solve the system of equations:

conservation of energy

( ) 0d dTY v V Hk k kdx dxkρ λ

⎛ ⎞⎜ ⎟+ − =⎜ ⎟⎝ ⎠∑ , (1.18)

where Hk the specific enthalpy of species k and λ the thermal conductivity coefficient.

With the proper choice of the boundary conditions for one-dimensional flames, it is

possible to solve the governing equations [2,16].

Various software packages have been developed, which are able to calculate the

one-dimensional flame structure in only a few minutes by solving the set of governing

equations. The simulation program used in this study is the PREMIX code [17]. This

code is included in the CHEMKIN II simulation package [13]. This package operates

using a reaction mechanism data file as input, along with thermal and transport

properties of the species involved in the mechanism. The program is able to calculate

temperature- and mole fraction profiles in both burner-stabilized and free flames.

1.4 Chemical mechanisms

In the last decades chemical kinetic mechanisms have been developed to model

combustion of hydrogen (for example, see [18]) and hydrocarbon mixtures ([11,19],

among many others). These mechanisms, used to describe the transformation of

reactants into products, may contain hundreds of species and thousands of elementary

reactions. Improvement of the mechanisms currently in use is necessary, since none of

them can be regarded as comprehensive [20], i.e. accounting for all combustion

phenomena and the predictive power is only accurate for a small range of parameters.

In order to improve the existing chemical mechanisms, they should be validated

against experimental data, where parameters are varied in a well-defined manner.

Sensitivity and rate-of-production analyses are used to design and optimize models.

Using these methods rate-limiting steps and characteristic reaction paths can be

identified [2].

Page 28: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

23

The experimental data obtained in this study have been modeled using different

mechanisms. One of these mechanisms is GRI-Mech 3.0 [11], which is widely used

and has arguably become the “industry standard” for methane in the research

community. This mechanism is optimized to model natural gas combustion and

contains 325 reactions and 53 species, including reactions that describe NO formation

and reburn chemistry.

Page 29: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

24

Literature

1. G. Dixon-Lewis., D. J. Williams., Comprehensive Chem. Kin. 17, (1977).

2. J. Warnatz, U. Maas, R. W. Dibble, Combustion, (Springer, Berlin, 1996).

3. C.-L. Yu, M. Frenklach, D. A. Masten, R. K. Hanson, C. T. Bowman, J.

Phys. Chem. 98 (1994) 4770-4771.

4. M. Frenklach, H. Wang, M. J. Rabinowitz, Prog., Energy Combust. Sci. 18:

(1992) 47-73.

5. B. Lewis, Q. von Elbe, Combustion Flames and Explosions of Gases, (Third

edition 1987).

6. C. K.Westbrook, F. L. Dryer, Prog. Energy. Combust. Sci. 10 (1984) 1-57.

7. J. M. Rabinowitz, J. W. Sutherland, P. M. Patterson, R. B. Klemm, J. Phys.

Chem. 95 (1991) 67-681.

8. B. E. Milton, J. C. Keck, Combust. Flame 58 (1984) 13-22.

9. C. K.Westbrook., Proc. of the Combust. Inst. 28 (2000) 1563-1577.

10. J. Huang, P. G. Hill, W. K.Bushe, S. R. Munshi, Combust. Flame 136 (2004)

25-42.

11. G. P. Smith, D. M. Golden, M. Frenklach, N. W. Moriarty, B. Eiteneer, M.

Goldenberg, C. T. Bowman, R. K. Hanson, S. Song, W.C. Gardiner, V.

Lissanski, Z. Qin, http://www.me.berkeley.edu/gri_mech/.

12. C. P. Fenimore, Proc. Combust. Inst. 13 (1971) 373-379.

13. A. E. Lutz, R. J. Kee, J. A. Miller, SENKIN: A FORTRAN program for

predicting homogeneous gas phase chemical kinetics with sensitivity

analysis. Sandia Report SAND87-8248. Sandia National Laboratories,

(1987).

14. R.J. Kee, F.M. Rupley, J.A. Miller, CHEMKIN II: A Fortran Chemical

Kinetics Package for the Analysis of Gas-Phase Chemical Kinetics., Sandia

National Laboratories, (1989).

15. R. M. Fristrom and A. A. Westenberg, Flame Structure, (McGraw-Hill, New

York, 1965).

16. R. J. Kee, F. M. Rupley, J. A. Miller, M. E. Coltrin, J. F. Grcar, E. Meeks, H.

K. Moffat, A. E. Lutz, G. Dixon-Lewis, M. D. Smooke, J. Warnatz, G. H.

Evans, R. S. Larson, R. E. Mitchell, L. R. Petzold, W. C. Reynolds,

Page 30: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 1

25

M. Caracotsios, W. E. Stewart, P. Glarborg, C. Wang, and O. Adigun,

CHEMKIN Collection, Release 3.6, Reaction Design, Inc., San Diego, CA,

(2000).

17. R. J. Kee, J. F. Grcar, M. D. Smooke, J. A. Miller, Fortran program for

modelling steady one-dimensional premixed flames. Sandia Report SAND85-

8240. Sandia National Laboratories, (1985).

18. O. M. Conaire, H. J. Curran, J. M. Simmy, W. J. Pitz, C. K. Westbrook, Int.

J. Chem. Kin. 36 (2004) 603-622.

19. http://www.chem.leeds.ac.uk/Combustion/Combustion.html

20. J. M. Simmie, Prog. Energy Combust. Sci. 29 (2003) 599–634.

Page 31: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

26

Chapter 2

The Rapid Compression Machine

Experimental Techniques, Procedures and Setup

Page 32: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

27

2.1 Background

Over the years, several facilities have been used to investigate autoignition under

strictly controlled experimental conditions, including flow reactors, shock tubes and

rapid compression machines (RCM). While each of these facilities has its merits, their

utility is restricted to certain ranges of pressure, temperature and ignition time.

Flow reactors: In a flow reactor, fuel is injected into a flowing air stream at high

temperature and/or pressure. The combustible mixture propagates through the reactor

and, depending on the velocity ignites at some distance downstream the fuel injector

location. Because the reaction zone is spread over a large distance, the flow reactor

offers the advantage of relatively simple measurements of the evolution of species

concentrations during the ignition process. The main drawback is that pressures

achievable in flow reactors are relatively low; further, since flow reactors makes use

of electric heaters, the maximum air temperature is limited, on the order of 1000K.

One of the most advanced flow reactors was developed at Princeton University, and

provides pressures up to ∼20 bar and temperatures up to ∼1200K [1,2].

Shock tubes: A shock tube uses the compressive heating of a shock wave to

bring a premixed combustible mixture to high temperature and pressure in a very

short time. A shock tube is ideal for studying ignition phenomena with short

characteristic times (order of tens of microseconds) under the conditions obtained. A

limitation of this technique is that the well-controlled test conditions persist for less

than 5 ms [3].

Rapid Compression Machines (RCM): The operating principle of the RCM is to

compress a homogeneous fuel/oxidizer mixture to moderate temperatures

(Tmax ≈ 1200K) and high pressures (Pmax ≈ 70bar) in a cylinder by the motion of a

piston. The RCM offers the advantage that the temperature and pressure of the

compressed mixture can be sustained for times longer than 10ms [4]. Moreover, it

provides a simple method of simulating the processes that take place in practical

devices such as spark engines and Homogeneous Charge Compression Ignition

(HCCI) engines. The time needed to compress the test gas mixtures limits the

minimum characteristic time of investigation to ∼1ms.

Several rapid compression machines have been developed and used to study

autoignition. The RCM developed at the University of Science and Technology at

Page 33: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

28

Lille is a right angle dual piston design RCM [5]. One of the pistons is air driven and

is connected by way of a cam to the other piston that compresses the mixture. The

cam controls the length of the stroke, the initial and final position of the compressing

piston, and prevents piston rebound after ignition. The maximum compression ratio

achievable with this machine is 10. Maximum pressures and temperatures after

compression are reported around 17 bar and 900K, with total times of compression of

20-60 ms. Minetti et al. used this RCM to study autoignition and two-stage ignition of

several hydrocarbon fuels [6-8]. In addition, this group performed measurements of

the temperature distribution in their RCM [4] and found that the gas temperature is

homogeneous for ∼15ms after compression, which is then distorted due to heat loss to

the wall. Griffiths et al. (University of Leeds) studied autoignition behavior of several

fuels [9-12] using an RCM that consists of a pneumatically driven piston. In support

of their experimental work they examined the development of the temperature field in

the combustion chamber of their machine [13,14] and observed that piston motion

during compression causes a roll-up vortex that moves “cold” gas from the wall into

the core. This resulted in a region with adiabatically heated gas directly after

compression containing a plug of colder gas. The Leeds RCM has a maximum

compression ratio around 15 and is able to compress the mixture in 22 ms. Final

pressures up to 20 bar and temperatures up to 1000K have been reported in this

machine. Park and Keck (MIT) developed an RCM [15,16] that consists of a

hydraulically operated piston-cylinder assembly. They also used a piston head with a

special crevice, designed to capture vortices created during compression [15,16], to

improve the homogeneity of the core gas. Lee and Hochgreb (MIT) optimized this

piston design for the suppression of the vortices [17,18], using results of detailed

modeling. Compression ratios of ∼19, maximum peak temperatures of ∼1200K and

maximum peak pressures of ∼70 bar can be achieved in this RCM. The gas mixture

can be compressed within 10 to 30 ms. Several autoignition experiments have been

performed with this machine [19,20]. Simmie and coworkers (National University of

Ireland, Galway) used an RCM, originally built at Shell laboratories [21] that uses two

horizontally opposed pneumatically driven pistons to rapidly compress the gas

mixture. For this machine, the maximum compression ratio reported is 13, the

compression time is ∼10 ms, the maximum peak compression pressure is 44 bar and

temperatures up to 1060 K have been reported. The machine has been used to study

Page 34: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

29

methane ignition [22], among other fuels. Also, this group confirmed the importance

of the crevice in the piston head [22,23]. Recently, a free-piston RCM had been

developed by Donovan et al. (University of Michigan). Compression ratios between

16 and 37, peak pressures around 20 bar and peak temperatures of 2000K have been

reported [24] using this RCM.

Given the wide range of pressures and temperatures achievable, we chose the

MIT design to study autoignition of CH4/H2/oxidizer mixtures. For this reasons a

replica of the MIT RCM was built in our laboratory and used to study autoignition.

Page 35: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

30

2.2 Design and Operation

Conceptually, the rapid compression machine simulates a single compression

event of an internal combustion engine. It is designed to compress a gas mixture in a

short time to high temperature and pressure while maintaining a well-defined uniform

core temperature in the reaction chamber (adiabatic core). Fast compression is

necessary to prevent substantial heat losses and radical build up before the end of the

compression. The piston is driven pneumatically and is decelerated smoothly by a

special hydraulic damper to reduce the impact velocity at the end of the compression.

The machine contains only two moving parts, the “fast acting valve” and the piston.

Both are made of aluminum, while all other parts are made from steel. The piston

used is hollow, to reduce its mass and to have uniformly distributed stress throughout

the body [17]. The piston head is removable and can be replaced by heads with other

crevice configuration.

Figure 2.1. Sketch of the cross sectional view of the RCM.

Page 36: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

31

The RCM, shown schematically in figure 2.1, and in detail in Appendix A.1,

includes a nitrogen-filled driving chamber, a speed-control oil chamber, an oil

reservoir chamber, a fast acting valve, a combustion chamber and a piston. The speed-

control oil chamber and the oil reservoir chamber, alternately connected and separated

by the fast acting valve are part of the hydraulic system that controls the movement of

the piston. The detailed sequence of operation for the RCM is given in Appendix A.2.

Briefly, after the RCM is triggered, the piston is in the “down” position and the fast

acing valve is in the “up” position, thus connecting the speed control oil chamber and

the oil reservoir chamber. To prepare for the next run, the piston is moved up until it

hits the stroke stop by pressurizing the oil reservoir chamber with ∼3 bar nitrogen.

The fast acting valve is than moved down by pressurizing the chamber above the fast

acting valve with 7 bar nitrogen, and locked in down position using 70 bar oil

pressure. The speed control and the oil reservoir chamber are now no longer

connected hydraulically. By pressurizing the speed control oil chamber with ∼48 bar

high-pressure oil, the piston is firmly locked in place against the stroke stop. After

loading the combustion chamber with the test gas mixture, the driver chamber is

pressurized with ∼35 bar nitrogen. The force on the piston created by the 35 bar

nitrogen pressure in the driving chamber is lower than the opposing force of oil on the

hydraulic piston, and hence the piston assembly is held in position by the stroke stop.

By opening the solenoid valve, the 70 bar oil pressure on the fast acting valve is

released and the fast acting valve will be pushed up by the 48 bar oil pressure in the

speed control chamber. The forces between the driving chamber and speed control

chamber are no longer balanced, and the pressure in the driving chamber causes the

piston to accelerate downward, compressing the test gas in the combustion chamber.

Subsequently, the piston’s acceleration slows, and the piston moves with constant

velocity until it is smoothly decelerated by a hydraulic damper [16]. In the final stage,

the deceleration force and velocity are reduced to zero, so that the final stop of the

piston at the bottom plate occurs without rebound. The piston is held firmly by the

force of driving nitrogen, which is greater than the force of the compressed gas

mixture in the reaction chamber. This allows combustion to take place at constant

volume. Since the area ratio of the piston on the driving side compared to the side of

combustion chamber is 4:1, the pressure inside the combustion chamber may be a

factor of 4 larger than the maximum pressure in the driving chamber before the piston

Page 37: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

32

will move. In the present construction, gas mixtures can be compressed with total

compression times of 15-30 ms up to pressures around 70bar. The piston speed can be

controlled, by varying the pressure in the driving chamber. The characteristics of the

RCM are presented in Table 2.1.

To cover a wide range of compression ratios, the rapid compression machine

was designed with adjustable piston stroke and clearance height. The piston stroke,

which is determined by the initial position of the piston, can be varied by turning the

stroke adjustment screw, see appendix A.1.. The clearance height can be changed by

replacing the clearance ring in the combustion chamber. To simulate temperatures and

pressures realistic for gas engines, different combinations of compression ratios,

initial pressures of the test gas and heat capacity of the diluent gases are used in this

study.

Table 2.1 RCM Characteristics

Cylinder bore 50.8 mm

Maximum stroke ∼160 mm

Maximum compression ratio ∼25

Clearance height ∼6.2-13 mm

Piston length ∼172 mm

Maximum driving pressure ∼35 bar

Maximum compression pressure ∼70 bar

Compression time ∼10-30 ms

2.2.1 Experimental System

Appendix A.3 shows the overall diagram of the RCM experimental system,

containing all main lines, pressure meters, oil drums, valves and the oil reservoir. All

lines are made from stainless steel with an inner diameter of 11 mm. The orifice

diameter of the solenoid valve has the same inner diameter (11 mm) as the lines, to

allow maximum throughput of oil. The high-pressure oil was supplied to the speed

control chamber (∼48 bar) and to the fast acting valve (∼70 bar) from two high-

Page 38: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

33

pressure oil accumulators. These oil accumulators contain oil and a bladder filled with

nitrogen to prevent mixing of nitrogen with oil. After ∼25 runs the accumulators were

refilled by oil from the main oil reservoir. Compressed nitrogen from five fifty-liter,

200 bar nitrogen bottles provided the required pressures in all parts of the system. The

operation pressures used, recommended by Park [16], are given in Appendix A.4

2.2.2 Gas filling system and filling procedure

All test gas mixtures were prepared in advance in a 10-liter gas bottle and used

to charge the combustion chamber at the required pressure. The gas filling system is

shown in figure 2.2. Before preparing the gas mixture, the gas bottle and the gas lines

connected to it are evacuated to less than 0.5 mbar using a vacuum pump. After

adding an individual component the bottle is closed. Subsequently, the gas lines are

again evacuated and the next mixture component is added to the bottle. The test

mixtures thus prepared were allowed to mix ∼24 hours to ensure homogeneity. Before

filling the combustion chamber, the poppet valve and the solenoid valve are opened,

and the whole system is evacuated to a pressure below 0.5 mbar. The combustion

chamber was filled with the gas mixture to the desired initial pressure, by opening the

bottle that contains the test gas. The poppet valve is then closed and the mixture is

ready for compression. After each run, the compressed gases in the RCM were vented

to the outside air, and the chamber was evacuated again before preparing the next run.

The solenoid valve in the gas-filling line was included for safety purposes, and

electrically connected such that when the solenoid valve used to trigger the fast-acting

valve is open, the solenoid in the gas-filling line is always closed. This prevents flame

propagation back to the gas-mixture bottle when the poppet valve is not properly

closed.

All test gases used in this study have purity greater than 99.5%. The composition

ratios of the gas mixtures are calculated from the measured partial pressures of the

individual gases.

Page 39: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

34

Figure 2.2. Gas handling system

2.2.3 Instrumentation and data acquisition

An MKS baratron diaphragm pressure gauge (type 722A) was used for

measuring all partial pressures of the components in the test mixture, and all other

pressures in the gas filling system. This pressure meter has an operating range from 0-

1300 mbar with an accuracy of 0.5% of reading. The dynamic pressures in the

combustion chamber during compression and throughout the post-compression period

were measured using a Kistler 6025B piezoelectric pressure transducer (range 1-250

bar, linearity ± 0.1%) placed at the bottom of the combustion chamber. The signal

from the transducer was amplified by a 5010B Kistler charge amplifier, recorded

digitally by an oscilloscope with a sample rate of 500 kHz and 16-bit resolution, and

processed by a PC. The initial temperatures of the mixture were measured by a Pt-Rh

thermocouple with an accuracy of ± 0.2K, located at the wall of the combustion

chamber. The data acquisition was triggered simultaneously with the opening of the

solenoid for the fast-acting valve.

Page 40: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

35

2.2.4. Determination autoignition delay time

A typical measured pressure trace is presented in figure 2.3. The gas mixture is

compressed in ∼20ms, to a peak pressure that indicates the end of the compression

event. The majority of the pressure rise in the compression period takes place in a

very short time (<3 ms). During this rapid compression heat losses and radical built up

are not substantial. After the peak compression pressure is reached, the pressure drops

gradually due to heat transfer to the walls. Subsequently, heat release due to

exothermic reactions causes a slight increase in pressure, followed by a sharp increase

in pressure indicating ignition.

Figure 2.3. Typical measured pressure trace for a stoichiometric CH4/H2//O2/N2/Ar mixture, with the definitions of autoignition delay time, peak pressure and compression time. The dotted line shows the calculated adiabatic peak pressure for the given compression ratio.

The auto ignition delay time is defined in this study as the time interval between

the peak pressure Pc that marks the end of compression and the time of maximum

pressure rise during ignition.

Page 41: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

36

2.2.5 Temperature Determination

To study autoignition behavior, one must know the instantaneous temperature in

the combustion chamber, since chemistry is very sensitive to temperature. However,

measuring the temperature in the reaction chamber directly by optical methods is

problematical, given the characteristic test times (order of a millisecond) and the

difficulty of making the reaction chamber optically accessible. The use of

thermocouples also permits in-situ temperature measurements, but the presence of the

thermocouple in the combustion chamber can significantly influence the

measurements, since the test gas can interact with the surface of the thermocouple.

For example, the surface of the thermocouple may act as a catalyst for chemical

reactions, and unintentionally induce ignition. The most straightforward method is to

calculate the temperature from the instantaneous measured pressure, assuming the

existence of an adiabatic core in the RCM chamber [18].

When an adiabatic system goes from one state (P1,T1,V1) to another state (P2,T2,V2),

initial and final parameters are related to each other by the isentropic relation (1.13) of

an ideal gas, which can be rewritten as,

∫ −=

2

12

1 ln1

1)ln(T

T

TdVV

γ, (2.1)

1

22

1

lnln1

1PP

TdT

T

=−∫ γ

, (2.2)

where γ(T) is the ratio of temperature-dependent heat capacities of the mixture at

constant pressure and constant volume, γ(T)=Cp(T)/Cv(T). The heat capacities used in

this study are taken from [25]. In figure 2.4, the temperature dependence of γ for two

inert gases, N2 and Ar, and the mixture H2/CH4/O2/Ar/N2 (0.95/0.05/1.85/3/4.3) used

in our experiments is illustrated (for a pressure trace for this mixture, see figure 2.3).

The figure shows that γ for Ar is much larger than that for both N2 and the

combustible mixture. Moreover, as expected for a monatomic gas, γ for Ar is constant

while those for N2 and the combustible mixture show a slight dependence upon the

temperature.

Page 42: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

37

Figure 2.4, Temperature dependence of γ for the gas mixtures N2, Ar and CH4/H2/O2/N2/Ar.

The relations (2.1) and (2.2) show that the pressure and temperature of the gas

only depend on the volumetric ratio and the specific heat capacities and the initial

conditions (Pi, Ti) of the gas mixture. Thus for an ideal RCM, the temperature (Tc)

and pressure (Pc) after compression can easily be calculated from equation (2.1) and

(2.2) by measuring only the initial temperature (Ti) and pressure (Pi), and the

mechanical compression ratio CR, which is defined as the ratio of the initial volume

(Vi) to the final volume (Vc) of the reaction chamber. As an example, the pressures

and temperatures are calculated, based on the mechanical compression ratios that can

be obtained in our RCM, for the three aforementioned gases/mixtures using the

relations (2.1) and (2.2), and presented in figure 2.5a and 2.5b. The calculations show

that the larger γ(T) for Ar, in comparison to N2 and the combustible mixture (figure

2.4), results in a much higher temperature and pressure after compression at identical

compression ratio. In reality, the temperatures and pressures will be lower than those

calculated due to heat losses during compression. As an example, for the compressed

CH4/H2/O2/N2/Ar mixture with CR=22.5 and Pi=0.49 bar, we measured a peak

pressure of ∼32 bar (figure 2.3). However, simple adiabatic calculations show that at

CR=22.5 the calculated temperature is 1050K (figure 2.5a) and Pc/Pi=80 (figure 2.5b),

this results in a calculated compressed pressure of ∼40 bar at Pi=0.49bar.

Page 43: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

38

Figure 2.5a) Temperature calculated as function of compression ratios for different gas mixtures at Ti=295K, using the relation (2.1). b) Pressure ratio (Pc/Pi) calculated as function of temperature for different gas mixtures at initial temperature Ti=295K, using the relation (2.2).

It is more realistic to make the assumption of the existence of an isentropically

compressed core region [11] that is unaffected by heat and mass transfer. In this case,

the relations (2.1) and (2.2) are valid for the pressures, temperatures and volumes in

the adiabatic core. To avoid the difficulty of calculating an effective compression ratio

based on the unknown volume of the core gas within the combustion chamber, the

ratio of measured pressures is used to calculate the temperature of the adiabatic core

gas using equation (2.2). The uncertainty of the calculated core gas temperatures (Tc)

is less than ±3.5K for all measurements. (Appendix A.5).

Page 44: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

39

Several studies [26,17] have indicated that vortices created by the motion of the

piston causes unwanted mixing of cold boundary-layer gas into the compressed core

gas, destroying the adiabatic core. It has been shown, both numerically [19,23,27]

using CFD calculations and experimentally [27] by temperature mapping using the

planar laser-induced fluorescence of acetone, that the incorporation of a specially

designed crevice on the piston head successfully suppresses the vortex formation, and

preserves the well-defined homogeneous core region intact. In this study, the creviced

piston head based upon the best design from the MIT RCM [17,18] was used.

Page 45: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

40

Appendix A.1

Figure A.1. Cross Sectional view of the Rapid Compression machine (RCM).

Driving chamber

Speed control chamber

Piston lock chamber

Fast acting valve

Stroke adjustment

Piston

Oil reservoir chamber

Poppet valve

Pressure transducer

Combustion chamber

Page 46: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

41

Appendix A.2

The numbers in the operation sequence table below are referring to the numbered

parts (valves, reducers etc.) in appendix A.3.

Step Operation Manometer Pressure (bar) Result

1 Close valve 1 2 Open valve 2 3 Open Valve 5 4 Reducer R2 M4 2 Venting air out of

chambers and lines 5 Close valve 2 6 Reducer R2 M4 48 Pressurizing oil drum 7 Close valve 4 8 Open valve 7 9 Reducer R5 M10 70 Pressurizing oil drum 10 3-way valve 4A in up

position Connected to solenoid

11 3-way valve 1A in up position

M2 0 Venting driving chamber

12 3-way valve 2A in down position

13 Reducer R3 M5/M8 1 3 Lifting piston 14 3-way valve 3A in

down position

15 Reducer R4 M9/M3 7 Lowering fast acting valve with N2

16 Open valve 4 17 Open solenoid M7 70 Locking fast acting

valve 18 Open valve 2 M1 48 Locking piston 19 3-way valve 3A in up

position M3 0 Venting N2 fast acting

valve 20 3-way valve 2A in up

position M5 0 Drain oil/N2

21 Open poppet valve 22 3-way valve 1A in

down position

23 Reducer R1 M6 30 Pressurizing 24 Close poppet valve 25 Close solenoid S1

(close = no current)

26 Close valve 4 27 3-way valve 4A in

down position

28 Check amplifier Reset 29 Close valve 2 30 Open solenoid S1 Fire the RCM

Page 47: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

42

Appendix A.3

Page 48: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

43

Appendix A.4

Table 3.4 Operating pressures Operation System Operating Pressure (bar)

Piston up Pneumatic 3.5

FAV*) down Pneumatic 7

Lock FAV Hydraulic 70

Lock piston Hydraulic 48

Driving pressure Pneumatic 35

*) Fast-acting valve

Page 49: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

44

Appendix A.5 Uncertainty analyses

As described above, the peak temperatures of the core gas at the end of

compression Tc were calculated from the equation (2.2). Since ignition chemistry is

very sensitive to temperature, it is important to estimate the uncertainty in the

calculated peak temperature caused by the uncertainties in the measured parameters.

Assuming that Ti, Pi, Pc and γ are uncorrelated, then the uncertainty in the calculated

peak temperature (ΔTc) can be determined from the following equation:

2T T T TT T P Pc i i cT P Pi i c

γγ

⎛ ⎞∂ ∂ ∂ ∂Δ = Δ + Δ + Δ + Δ⎜ ⎟∂ ∂ ∂ ∂⎝ ⎠

, (A.2)

As mentioned in 2.2.3,

ΔTi=±0.2K

ΔPi/P=0.5%

ΔPc/P=0.1%

and Δγ=0.04, based on the accuracy of the determined Cp values [25]. Figure A.2

shows the calculated uncertainty as function of the peak pressure (Pc), by using

equation A.2.

Page 50: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

45

Figure A.2. The calculated uncertainty in the peak temperature.

The figure shows that the uncertainty in the peak temperature is better than ±3.5K in

the range of pressures of interest (10-70 bar).

Page 51: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

46

Literature

1. M. L. Vermeersch, T. J. Held, Y. Stein, F. L. Dryer, SAE paper, 912316

(1991).

2. T. J. Kim, R. A. Yetter, F. L. Dryer, Proc. Combust. Inst. 25 (1994) 759-766.

3. A. G. Gaydon, I. R. Hurle, The Shock Tube in High-Temperature Chemical

Physics, Reinhold, New York, 1963.

4. P. Desgroux, R. Minetti, L. R. Sochet, Combust. Sci. Technol. 113 (1996)

193-203.

5. M. Carlier, C. Corre, R. Minetti, J.F. Pauwels, M. Ribacour, L. F. Socket,

Proc. Combust. Inst. 23 (1990) 1753-1758.

6. R. Minetti, M. Ribaucour, M. Carlier, L.R. Sochet, Combust. Sci. Technol.

113-114 (1996) 179-192.

7. R. Minetti, M. Carlier, M. Ribaucour, E. Therssen, L. R. Sochet, Combust.

Flame. 102 (1995) 298-309.

8. R. Minetti, M. Carlier, M. Ribaucour, E. Therssen, L. R. Sochet, Proc.

Combust. Inst. 26 (1996) 747-753.

9. J. F. Griffiths, P. A. Halford-Maw, D.J. Rose, Combust. Flame. 95 (1993)

291-306.

10. P. Beeley, J. F. Griffiths, P. Gray, Combus.t Flame 39 (19980) 269-281.

11. J. F. Griffiths, Q. Jiao, W. Kordylewski, M. Schreiber, J. Meyer, K. F.

Knoche, Combust. Flame, 93 (1993) 303-315.

12. A. Cox, J. F. Grifiths, C. Mohamed, H.J. Curran, W. J. Pitz, C. K.

Westbrook, Proc. Combust. Inst. 26 (1996) 2685-2692.

13. J. Clarkson, J.F. Grifiths, J. P. Macnamara, B. J. Whitaker, Combust. Flame

125 (2001) 1162-1175.

14. J.F. Grifiths, J.P. Macnamara, C. Mohamed, B.J. Whitaker, J. Pan, C.G.W.

Sheppard, Faraday Discussion, 119 (2001) 287-303.

15. P. Park, J. C. Keck, SAE Paper 900027 (1990).

16. P. Park, ‘Rapid Compression Machine Measurements of Ignition Delays for

Primary Reference Fuels’, Ph.D. Thesis, MIT, 1990.

17. D. Lee, S. Hochgreb, Combust. Flame, 114 (1998) 531-545.

18. D. Lee, ‘Autoignition Measurements and Modeling in a Rapid Compression

Machine’, Ph.D. Thesis, MIT, 1997.

Page 52: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 2

47

19. D. Lee, S. Hochgreb, J.C. Keck, SAE paper 932755 (1993).

20. D. Lee, S. Hochgreb, Int. J. Chem. Kin. 30 (1998) 385-406.

21. W. S. Affleck, A. Thomas, Proc. Inst. Mech. Eng. 183 (1969) 365-385.

22. L. Brett, J. Macnamara, P. Musch, J. M. Simmie, Combust. Flame, 124

(2001) 326-329.

23. J. Wurmel, J. M. Simmie, Combust. Flame, 141 (2005) 417-430.

24. M. T. Donovan, X. He, B. T. Zigler, T. R. Palmer, M. S. Wooldridge, A.

Atreya, Combust. Flame 137 (2004) 351-365.

25. B. J. McBride, S. Gordon, M. A. Reno, 'Coefficients for Calculating

Thermodynamic and Transport Properties of Individual Species', NASA

Report TM-4513, October 1993

26. R. J. Tabaczynski, D. P. Hoult, J.C. Keck, J. Fluid Mech., 42 (1970) 249-

255.

27. G. Mittal, C. J. Sung, Combust. Flame 145 (2006) 160-180.

Page 53: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

48

Chapter 3

High-pressure autoignition delay time measurements in methane/hydrogen fuel mixtures in a Rapid Compression Machine

Page 54: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

49

3.1 Introduction

Increasingly stringent regulations regarding CO2 emissions, and the knowledge

that fossil fuel reserves will be exhausted within this century, have called attention to

the possible use of admixtures of hydrogen in natural gas as an alternative fuel in

combustion devices. Experimental results [1] have shown that addition of small

amounts of hydrogen to methane, the principal component of natural gas, enhance the

performance of a gas-powered spark-ignited engine. In the same paper, numerical

simulations were used to indicate that hydrogen addition to methane significantly

increases the tendency to knock at hydrogen fractions larger than 20%. Knocking

combustion in spark-ignited engines is closely related to autoignition of the unburned

end gas, and should be avoided at all cost since it can physically damage the engine

and increase pollutant emissions. Of course, the methane number [2], used to

characterize the knock tendency of natural gases, takes the difference in knock

behavior between methane and hydrogen to define the extremes of the scale, with

pure methane as most knock resistant and pure hydrogen as the least resistant.

Understanding autoignition behavior is also important for designing gas turbines [3]

and Homogeneous Charge Compression Ignition (HCCI) engines [4] Moreover,

autoignition delay times are used as targets for the development and benchmarking of

chemical kinetic models for combustion.

While a large number of studies of the ignition of methane and hydrogen have

been reported, most of them have been conducted under diluted conditions (fuel mole

fraction < 10%) at relatively low pressures (< 5bar). Autoignition studies under

conditions relevant to engines are scarce. Undiluted H2/O2/Ar/N2 [5] and H2/O2/Ar [6]

mixtures have been studied in a rapid compression machine (RCM) at temperatures

ranging from 950 to 1100K and pressures lower than 50 bar. Shock tube

measurements of the autoignition delay times in slightly diluted CH4/O2 mixtures

(fuel + oxidizer ∼30%) at high pressures (40-240 bar) and intermediate temperatures

(1040 - 1500 K) have also been reported [7]. Ignition delay times in CH4/O2/Ar

mixtures using an RCM [8] have been measured at 16 bar between 980 and 1060K.

Ignition delay times obtained in non-diluted lean (ϕ=0.5) methane/air mixtures behind

reflected shock waves between 3 and 450 bar and at temperatures from 1300 to

1700 K [9] were compared with those calculated using the GRI-Mech 3.0 chemical

mechanism [10] and showed good agreement. To our knowledge, only three studies of

Page 55: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

50

autoignition in hydrogen/methane fuel mixtures have been reported [11-13]. In [11]

the influence of small additions of hydrogen (2 and 15% of the fuel by volume) to

highly diluted methane/air mixtures at high temperatures (1500 – 2150 K), moderate

pressures (2 - 10 bar) and equivalence ratios ranging from 0.5 to 2.0 was studied using

shock tubes. A thermal-based promotion theory was proposed to account for the effect

of hydrogen addition. A very extensive shock tube study of hydrogen/methane

mixtures (temperatures and pressures ranging from 800 to 2000 K and from 1 to 3 bar,

respectively) has been reported in [12]. The ignition delay time τ of dilute H2/CH4

mixtures was related to the ignition delay times of pure gases through the empirical

relation

(1 )4 2CH Hβ βτ τ τ−= , (3.1)

where τH2 and τCH4

are the ignition delay times of hydrogen and methane,

respectively, and β is the mole fraction of hydrogen in the fuel. Recently, the

autoignition delay times of two stoichiometric CH4/H2/air mixtures at pressures from

16 to 40 atm and temperatures between 1000 and 1300 K have been measured in a

shock tube [13]. Because the well-controlled test conditions in a shock tube persist

only for a few milliseconds, the combination of pressure and temperature in this study

was chosen to give ignition delay times <3 ms. Interestingly, while a relatively large

amount of hydrogen was added to the fuel (35%), only a relatively small reduction in

ignition delay time was observed compared to that observed for pure methane. The

experiments also show that the ignition-enhancing effect of hydrogen decreases with

decreasing mixture temperature, and decreases significantly upon increasing pressure

from 16 to 40 atm. The experimental results were compared with calculations

performed using a mechanism [14] that was a modified version of that taken from [7];

the comparison showed substantial disagreement, prompting the authors [13] to

recommend additional experimental and kinetic studies aimed at the autoignition

behavior of methane/hydrogen mixtures.

In this Chapter, we report the autoignition delay times of stoichiometric

methane/hydrogen mixtures using oxygen/nitrogen/argon oxidizers at high pressures

(10 – 70 bar), and temperatures from 950 to 1060 K. The pressures and temperatures

Page 56: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

51

of the unburned mixtures were chosen to give ignition delay times ranging from 2 to

50 ms. The measurements were performed in an RCM and compared with numerical

simulations using different chemical mechanisms, taking into account heat loss

occurring in the period between compression and ignition.

3.2. Experimental approach

The measurements have been performed in an RCM. The RCM construction,

operating procedure, gas handling system and pressure measurements are described in

detail in Chapter 2. The H2/CH4 mixtures containing 0, 5, 10, 20, 50 and 100% of

hydrogen are used as fuel. The compositions of gas mixtures used are given in

Table 3.1.

Table 3.1

Compositions of mixtures used in ignition experiments

Mixture [H2] [CH4] [O2] [N2] [Ar]

A 1 0 0.5 0 2.5

B 1 0 0.5 1.05 0.95

C 0.5 0.5 1.25 2.18 2.83

D 0.2 0.8 1.7 2.85 3.95

E 0.1 0.9 1.85 3.07 4.34

F 0.05 0.95 1.93 3.18 4.53

G 0 1 2 3.3 4.7

H 0.5 0.5 2.49 3.36 6.6

The fuel and oxygen concentrations in the mixtures A-G are in stoichiometric

proportions and mixture H is a fuel lean mixture (ϕ=0.5). The total concentration of

diluting inert gases are close to that of nitrogen in air, while the N2/Ar ratio is chosen

to provide similar temperatures after compression for all fuels. For comparison with

the results of a previous RCM study [6], the measurements are also performed for

pure hydrogen without nitrogen (flame A in Table 3.1). As mentioned above, the

pressures were varied between 10 and 70 bar, and temperatures in the range 950-

1060 K. In addition, a substantial number of measurements were performed along an

isotherm at a peak temperature after compression of 995 ± 4 K between ~25 and ~65

Page 57: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

52

bar (see below), allowing an examination of the pressure dependence of ignition. For

measurements under identical conditions (composition, initial/final pressure), the

reproducibility of the measured ignition delay times is ∼5% and the uncertainty in

deriving the ignition delay time from the measurements is ∼0.3ms. For the data taken

along the isotherm, the variation in temperature (± 4 K) causes a scatter in the results

of ~10-20%.

3.3. Numerical simulation and analysis of experimental data

3.3.1. Chemical mechanisms In this work we compare different chemical mechanisms for the calculation of

the ignition delay times. In the discussion of the mechanisms it will be convenient to

refer to them either by acronym (e.g., GRI-Mech) or by author. While one does not a

priori anticipate good performance from the GRI-Mech 3.0 chemical mechanism [10]

since it was optimized to model natural gas combustion over the ranges 1000 -

2500 K, 0.013 – 10 atm and equivalence ratios from 0.1 to 1.5, the large popularity of

this mechanism compels us to evaluate its predictive power under the experimental

conditions discussed here. Better predictions at high pressures can be expected from

the RAMEC mechanism[7], which includes 190-reactions involving 38 species based

on the GRI-Mech 1.2 mechanism [15], with additional reactions important in methane

oxidation at lower temperature. This mechanism emerged from the kinetic study

based on the high-pressure shock tube measurements referred to above [7], and

showed the increased importance of reactions involving HO2, CH3O2 and H2O2 at high

pressures and low temperatures. The comprehensive chemical mechanism for

methane oxidation, developed at the University of Leeds [16], is also used in the

present study. This mechanism consists of 351 chemical reactions between 37 species,

and is built on the same experimental base as GRI-Mech 3.0. Recently, two revised

mechanisms for hydrogen oxidation have been reported. A mechanism for hydrogen

oxidation consisting of 19 reversible elementary reactions has been developed by

O’Connaire et al. [17] and evaluated for temperatures ranging from 298 to 2700K,

pressures from 0.05-87 bar and equivalence ratios in the range from 0.2-6. The

hydrogen mechanism developed by Li et al. [18] is close to that of O’Connaire et al.

Page 58: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

53

and also includes 19 chemical reactions. The mechanism of O’Connaire is included in

the comprehensive kinetic model of methane/propane oxidation of Petersen et al.[19],

which consists of 663 chemical reactions between 118 species. In this mechanism, the

methane oxidation chemistry incorporates recent theoretical and experimental data for

the reaction rates. Clearly, the assumption is that a mechanism that performs well for

both pure hydrogen and pure methane will adequately describe H2/CH4 mixtures as

well.

3.3.2. Numerical simulations As mentioned in Chapter 1, the system of governing equations describing the

time evolution of the adiabatic core is not closed. Therefore, for meaningful

comparison between measurements and numerical simulations, the experiment should

provide sufficient information to specify the system for the simulation. In RCM

experiments, pressure can be measured relatively easily with a high degree of

accuracy, and is generally used for this purpose. To our knowledge, all efforts thus far

have been directed to estimating the specific volume (equation 1.14) from the pressure

trace. As was mentioned in Chapter 2, estimating the specific volume of the adiabatic

core based on the geometrical size of the combustion chamber is very inaccurate. A

more accurate approach is to calculate the specific volume directly after compression

assuming that the gas mixture is chemically inert (no heat release) and neglecting heat

losses between compression and ignition. For ignition delay times that are longer than

the time of compression, neglecting heat release from chemical reactions is a

reasonable assumption during compression. However, significant cooling of the

compressed gas mixture before ignition (as illustrated in Figure 2.3) is expected under

typical RCM conditions. To overcome this difficulty, the system of governing

equations (1.6)-(1.9) is supplemented by additional equations taking into account heat

and mass exchange in the RCM [20,21]. Unfortunately, the complex geometry of the

RCM requires many assumptions, and the implementation of this approach ultimately

requires the use of parameters derived from fitting the pressure trace obtained in an

inert gas mixture. In an alternative approach [5], the specific volume is calculated

from the measured pressure trace in an inert mixture with the same values of heat

capacity, initial pressure and temperature as those in the reactive mixture under

investigation. Clearly, this method accurately predicts the specific volume of the

Page 59: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

54

reactive mixture during the initial stage of the ignition process, when heat release is

marginal. Just prior to ignition, the specific volume can be substantially different from

that of the inert mixture as result of the pressure and temperature increase in the

adiabatic core due to chemical reactions. The error arising from the neglect of heat

release in the estimate of the specific volume on the calculated ignition delay time is

small if the duration of this phase is also small. We anticipate that this effect will be

larger for complex alkanes showing multistage ignition [21] than for the simple fuels

used here, for which only negligible heat release occurs prior to ignition.

In light of these considerations, in the present work we determine the specific

volume from the measured pressure in the period between compression and ignition,

and extrapolate the time dependence derived in this fashion to the region in which

substantial heat release begins. This method bypasses time consuming measurements

in inert gas mixtures, and, as will be seen below, yields fits to the pressure trace that is

on par with those obtained by the other methods [5,21]. While our approach is also

not free from the potential uncertainties due to heat release during ignition, the good

fit to the pressure trace indicates an accuracy of the same order that obtained by

determining the specific volume in the inert mixture.

The implementation of the method is illustrated in figure 3.1, where the

measured pressure trace in mixture B (Table 3.1) is presented together with the

specific volume derived from the pressure. The pressure in the interval from tmax

(point during compression at which the pressure reaches its maximum, Pmax) to tA (at

which heat release is still negligible) is approximated by an exponential time

dependence with a characteristic time τc:

0max

0max/)(exp)()( PPPtP c +−−−= τττ , (3.2)

where P0 is the pressure in the combustion chamber at room temperature. The specific

volume, v(t), is calculated from this relation assuming adiabatic expansion of the

“inert” mixture in the core

,/1)()(

maxmax

γ

⎟⎟⎠

⎞⎜⎜⎝

⎛=

PtPvtv (3.3)

Page 60: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

55

where vmax is the specific volume at tmax, and γ is Cp/Cv; v(t) used as an input in

SENKIN for all times t > tmax. The choice of tA is important for the accuracy of this

method. If tA is chosen too large, heat release due to chemical reactions will be

sufficient to substantially slow the pressure drop by heat loss, and the specific volume

thus derived will be inaccurate. To avoid this error, we determine tA using an iterative

procedure. In first approximation, tA is chosen to lie halfway between tmax and the time

of ignition. If the SENKIN calculations using v(t) yield a temperature T(tA), which is

2 K higher than that for the adiabatic core calculated from the measured pressure

trace, tA is decreased (shifted towards tmax); if T(tA) is within 2 K, tA is increased. The

iterations continue until increasing tA increases the temperature by more than 2 K.

Usually, 3 – 4 iterations were performed when analyzing the experimental results. It

should be pointed out that this iteration procedure is coupled to the specific chemical

mechanism being used in the simulations, which itself is the subject of investigation.

Therefore, a computed ignition delay time that is too short by more than a factor of

two or three can result in a value for tA that is too short for a reliable extrapolation.

Fortunately, the chemical mechanisms used here predict the ignition delay with

sufficient accuracy (within a factor of two) to implement the iteration procedure, and

the temporal profiles of specific volume used in the simulations are independent of the

choice of tA.

Page 61: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

56

Figure 3.1. Pressure (a) and specific volume (b) traces in mixture B (Table 3.1) at initial temperature of 995 K. Curves 1, 2 and 3 denote the results from the measurements, the calculations assuming constant specific volume after compression, and the calculations using the specific volume derived from the measured pressure trace, respectively.

The simulated pressure trace using the mechanism of O’Connaire et al. [17] are

presented in figure 3.1. Here we see that the simulated trace agrees with the

experimental trace as well as that reported by the other methods [5,21]. For

comparison, the pressure trace calculated assuming negligible heat loss (specific

volume is constant after compression) is also shown. As can be seen, neglecting heat

loss in the RCM can lead to the (erroneous) conclusion that this chemical mechanism

underpredicts the ignition delay time. However, the more realistic input of an

increasing specific volume shows opposite result – the mechanism actually

overpredicts the ignition delay time.

3.4. Results and discussion

To assess the quality of the experimental data, we compare the ignition delay

times obtained here with the results of previous RCM studies of the autoignition of

pure hydrogen [5,6]. The ignition delay times from different data sets are scaled by

Page 62: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

57

the oxygen number density (mol/cm3) at the peak pressure after compression [6], and

are presented in figure 3.2 as a function of the reciprocal peak temperature after

compression. As mentioned above, the measured ignition delay time depends

substantially upon the heat losses in the RCM combustion chamber. Based upon the

simulations, we estimate that heat loss in our RCM results in as much as a 35%

increase in the ignition delay time as compared with an ideal adiabatic RCM. As can

be seen from figure 3.2, all experimental results are within an interval of ±35% of our

measurements. Taking into account the assumption of the validity of the scaling

method, and that heat loss can vary significantly between the physically different

machines, we consider the agreement of the results obtained here and the data in the

literature to be excellent.

Figure 3.2. Scaled ignition delay times in pure hydrogen fuel as a function of reciprocal temperature after compression. Solid lines denote ±35% interval around the measurements in present work approximated by the relation (3.4), see text. As can be seen from figure 3.2, the ignition delay time in the pure hydrogen fuel

is exponential function of the reciprocal temperature. The same dependence is

observed for pure methane. Incorporating the pressure dependence using the power

function for number density, we obtain an Arrhenius-like empirical relation for the

functional dependence of the ignition delay time upon pressure and temperature after

compression (Pc and Tc, respectively) for pure hydrogen and methane fuels:

Page 63: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

58

⎟⎟

⎜⎜

⎟⎟

⎜⎜

⎛=

cRTaEn

cTcP

A expτ . (3.4)

The magnitudes of A, n, and Ea (units: s, Pa, mole, kJ, K) derived for the

stoichiometric mixtures of hydrogen and methane with oxygen are given Table 3.2.

Table 3.2 Fit coefficients

A Ea n

H2 2.82E-13 336 -1.3

CH4 3.23E-2 192 -2.1

It should be pointed out that the negative value of the power n for both gases means

decreasing ignition delay time with increasing pressure. We remark in passing that the

apparent activation energy for hydrogen is in excellent agreement with that observed

in Ref. [6]; while for methane Ea is significantly lower than that obtained in recent

studies [7,14].

To assess whether the recommended mixing expression (3.1) can also be used

for methane/hydrogen mixtures studied in the RCM, ignition delay times have been

measured at constant peak pressure (Pc = 33.5 ± 1 bar) and peak temperature

(Tc = 995 ± 4 K) as a function of hydrogen mole fractions in the fuel. As can be seen

from the results, presented in figure 3.3, replacing methane by hydrogen decreases the

ignition delay time, as reported in shock tube studies [11-13]. Moreover, within the

limits of the experimental uncertainty, the logarithm of the ignition delay time appears

to be linear function of the hydrogen mole fraction, suggesting the utility of mixing

expression (3.1). Anticipating the discussion below, we note that the computed

ignition delay times, using the mechanism from Ref. [19] and accounting for heat loss

as described in Section 3.3.2, predict this trend within the limits of experimental error.

For further analysis, we rewrite the mixing expression (3.1) by using Equation (3.4)

for the ignition delay time in pure fuel:

Page 64: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

59

2 42 4

2 4

(1 ) (1 )(1 ). expH CH

H CHH CH

n n E EPcA AT RTc

β β β βββτ+ − + −⎛ ⎞ ⎛ ⎞−= ⎜ ⎟ ⎜ ⎟⎜ ⎟⎜ ⎟ ⎝ ⎠⎝ ⎠

. (3.5)

From this relation, at fixed hydrogen mole fraction we expect a linear dependence of

the logarithm of ignition time divided by number density to the power

n = nΗ2β + nCH4(1-β) upon the reciprocal of the temperature. As can be seen from

figure 3.4, which shows the results for different hydrogen mole fractions, the mixing

relation (3.5) approximates the experimental data very well. For the mixtures with H2

content ≤ 20% the effect of hydrogen addition on the ignition delay time is relatively

small, but becomes substantial when the hydrogen fraction is more than 50%. It is

interesting to note that the slope of the lines in figure 3.4 increases with increasing

hydrogen content in the mixture, reflecting the differences in the “overall” activation

energy Ea between the two pure fuels; for hydrogen Ea is two times larger than for

methane (Table 3.2). Thus, at high temperatures, effect of the added hydrogen on the

ignition delay time is more pronounced than at low temperatures, as also observed in

[13].

Figure 3.3. Measured and calculated ignition delay times versus hydrogen mole fraction in fuel at Pc = 33.5 ± 1 bar and Tc = 995 ± 3 K). The simulations were performed accounting for heat loss using the mechanism of Petersen et al. [19]. The solid line is obtained from the mixing relation Equation 3.5, see text.

Page 65: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

60

Figure 3.4. Measured ignition delay times scaled according to equation 3.4 as a function of reciprocal temperature (symbols), and the calculated results using the mixing relation (3.5) (lines).

As can be seen in figure 3.4, all measurements obtained along the isotherm at peak

compression temperature ~995 K and fixed hydrogen mole fraction collapse to a small

cluster, which demonstrates that equation (3.5) correctly predicts the pressure

dependence of the ignition delay time. Presenting the isotherm data in figure 3.5 on a

linear scale, as a function of pressure for different volume fractions of hydrogen, we

observe some scatter around the lines from equation (3.5), which is caused by day-to-

day variations in temperature (± 4 K) in the measurements, as mentioned in section

3.2. As observed above, the ignition delay times decrease with increasing pressure for

all hydrogen fractions measured, extending the observations of the recent shock tube

study [13].

Page 66: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

61

Figure 3.5. Measured (symbols) and calculated (lines) autoignition delay times as a function of pressure at fixed peak compression temperature Tc=995 ± 4 K and different hydrogen mole fractions in fuel. The calculated curves were obtained using mixing relation (3.5),

3.5. Comparison of experimental results with numerical simulations

Figures 3.6 to 3.8 show the experimental and calculated autoignition delay times

using different chemical mechanisms. To avoid clutter in the figures, the simulated

data are presented as polynomial trend lines through the calculated points. The

measured and calculated ignition delay times are presented in two sets. The first set

(figures 3.6a-3.8a) includes logarithms of the ignition delay times scaled by (P/T)n as

a function of the reciprocal temperature to eliminate the density dependence and

highlight the expected Arrhenius behavior with temperature. The second set of figures

presents the ignition delay times measured along the 995 ± 4 K compression isotherm.

Page 67: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

62

Figure 3.6. Measured (diamonds) and calculated (lines) autoignition delay times for pure hydrogen (mixture B in Table 3.1). (a) Scaled delay times vs. reciprocal temperature; (b) delay time vs. pressure at fixed temperature Tc = 995 ± 4 K.

The results for pure hydrogen (mixture B, Table 3.1) are presented in Fig 3.6. As

can be seen, the calculations using the mechanism from Petersen et al. show excellent

agreement with the measurements over the entire range of pressure and temperature

studied, while the mechanism of Li et al. and the Leeds and RAMEC mechanism

systematically overpredict the measured ignition delay times. Calculations using the

Page 68: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

63

Leeds mechanism and GRI-Mech 3.0 (not shown in figure 3.6) give identical results

for all experimental conditions, which considering their similarity is perhaps not

surprising. Based on the decidedly better agreement obtained using Petersen et al. we

suggest the use of this mechanism for ignition delay studies of hydrogen combustion

under gas turbine conditions, similar to the recommendation made in a recent study

[22].

Figure 3.7 presents the measured and calculated ignition delay times for pure

methane (mixture G, Table 3.1). As can be seen from figure 3.7a, the calculations

using the Leeds and Petersen et al. mechanisms are in excellent agreement with the

experimental results for all conditions measured. The predictions of the RAMEC

mechanism are in reasonable agreement with the experiments at the low temperatures

but substantially underpredict (up to a factor of two) the ignition delay times at high

temperatures. The results of the calculations with GRI-Mech 3.0 are more than a

factor of two higher than the scaled measured ignition delay times for all data in

figure 3.7a. The unscaled data for ignition delay times along the isotherm at

Tc = 995 K presented in figure 3.7b also show excellent agreement between

measurements and calculations with the Leeds and Petersen et al. mechanisms. At this

temperature the RAMEC mechanism slightly underpredicts the measurements at

pressures below 45 bar, but improves with increasing pressure. GRI-Mech 3.0, while

following the experimental trend well, substantially overpredicts the ignition delay in

this range.

Page 69: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

64

Figure 3.7. Measured (diamonds) and calculated (lines) ignition delay times for pure methane (mixture G in Table 3.1). (a) Scaled delay times vs. reciprocal temperature; (b) delay time vs. pressure at fixed temperature Tc = 995 ± 4 K.

As observed above, the mechanism proposed by Petersen et al. predicts the ignition

delay time in both pure hydrogen and methane very well, while the predictions of the

other mechanisms considered are poorer. Furthermore, this mechanism also yields the

best agreement with the experiments performed on the hydrogen/methane mixtures.

Consequently, in the following comparisons we only show these computational

results. As can be seen from figure 3.8, the computed ignition delay times are in

Page 70: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

65

excellent agreement with the experiments in the hydrogen/methane mixtures. Over the

entire range of pressure and temperature studied, the agreement between the

calculations and measurements is better than 25%.

Figure 3.8. Measured (symbols) and calculated (lines) autoignition delay times for hydrogen/methane fuel (mixtures D, E and F in Table 3.1) as a function of reciprocal temperature (a) and pressure at fixed temperature Tc = 995 ± 4 K (b). The calculations were performed using Petersen et al.[19].

Page 71: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

66

Given the agreement with the experimental results presented here, we also

compare the predictions of this mechanism with the experimental data reported in Ref.

[13], where a significant disparity between the experimental and numerical results

was observed. Figure 3.9 reproduces the experimental data for the hydrogen/methane

mixtures from Ref. [13], and the computations using Petersen et al.; as was done in

the original report [13], we have simulated the results at constant density. At 16 bar

(figure 3.9a), the current mechanism gives a better reproduction of the experimental

data at T > 1150K, the computations for 15 and 35% now bracketing the experiments;

in this region, the agreement is substantially better than a factor of 2. However, at

lower temperature the agreement is poorer than in the original model [13]. At 40 bar

(figure 3.9b), we observe a similar trend, albeit with a similar agreement to the

original model at lower temperatures. Particularly vexing is the relatively large effect

for the addition of hydrogen predicted by the computations, but which is apparently

much less manifest in the shock tube results, even at high temperature. We remark

that the good predictive power shown in figure 3.8, particularly in reproducing the

trends with pressure and hydrogen addition as well as the magnitude of the results in

the low temperature region, supports the use of the mechanism under these conditions.

Page 72: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

67

Figure 3.9. Comparison of measured (points) and calculated (lines) ignition delay times for shock tube measurements taken from Ref. [13]. Calculations performed using mechanism from Petersen et al. [19]; dashed lines calculations at 15% H2, solid lines calculations at 35% H2 in mixture. Figure (a) at 16 bar, figure (b) at 40 bar.

While the discussion above has focused entirely on stoichiometric mixtures, we

have also obtained results under lean conditions (ϕ = 0.5), for the 50/50

hydrogen/methane mixture. The composition used in these measurements is shown in

Table 3.1 as mixture H. The ignition delay times measured along the 995 K isotherm

are presented in figure 3.10, together with the stoichiometric results. Interestingly, we

see no change in the measured delay times between the lean and stoichiometric

mixtures within the experimental uncertainty (dominated by the ± 4 K temperature

uncertainty discussed above). In addition, as seen in the previous figures, the

computational results using Petersen et al. predict both trends and magnitude of the

results excellently. Under the conditions of the experiments the predicted differences

for the two equivalence ratios is less than 1 millisecond. We are currently extending

these measurements to other hydrogen/methane ratios.

Page 73: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

68

Figure 3.10. Measured (symbols) and calculated (lines) ignition delay times for 50% H2 in the fuel at ϕ = 1.0 and ϕ = 0.5 (Mixtures C and H, respectively, in Table 3.1) as

a function of pressure at fixed temperature Tc = 995 ± 4 K. The calculations were performed using Petersen et al. [19].

3.6 Summary and Conclusions

Autoignition delay times of methane/hydrogen mixtures at high pressure (10-

70 bar) and moderate temperatures (960 – 1060 K) have been measured in a rapid

compression machine. The experimental results obtained under stoichiometric

conditions show that replacing methane by hydrogen reduces the measured ignition

delay time. Both measured and computed ignition delay times in the fuel mixtures are

shown to be related quantitatively to the hydrogen mole fraction in fuel according to

the mixing relation proposed in the literature [12]. At low mole fractions (≤20%),

hydrogen addition has a modest effect on the measured ignition time under the

experimental conditions presented here. At 50% hydrogen mole fraction in fuel a

substantial reduction in ignition delay time is observed. The measurements show that

the effects of hydrogen in promoting ignition increases with temperature and

decreases with pressure. Interestingly, results for 50% hydrogen in the fuel at

ϕ = 0.5 are essentially identical to those at ϕ = 1.0. These results suggest that the

adverse affects of hydrogen addition to natural gas on engine knock may be limited

for hydrogen fractions of only a few tens of percent. Very good agreement between

the measurements and calculations using the mechanism proposed by Petersen et al.

Page 74: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

69

[19] is observed for all fuel mixtures studied. Over the entire operational range of

temperatures and pressures used in the present study, the differences between the

measured and calculated values of the ignition delay time are less than 10% for pure

fuels and better than 25% for the hydrogen/methane mixtures.

Page 75: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

70

Literature

1. G. A. Karim, Int. J. hydrogen energy 28 (5) (2003) 569-577.

2. M. Leiker, W. Cartelliere, H. Christoph, U. Pfeifer, M. Rankl, ASME paper

72-DGP-4, April 1972.

3. J. Y. Ren, F. N. Egolfopoulos and T. T. Tsotsis, Combust. Sci. Technol 174

(4) (2002) 181-205.

4. M. Richter, R. Collin, J. Nygren, M. Alden, L. Hildingsson and B.

Johansson, JSME Int. J. Series B-Fluids and Thermal Engineering 48 (4)

(2005) 701-707.

5. G. Mittal, C. J. Sung and R. A. Yetter, Int. J. Chem. Kin. 38 (8) (2006) 516-

529.

6. D. Lee and S. Hochgreb, Int. J. Chem. Kin. 30 (6) (1998) 385-406.

7. E. L. Petersen, D. F. Davidson and R. K. Hanson, Combust. Flame 117 (1-2)

(1999) 272-290.

8. L. Brett, J. MacNamara, P. Musch and J. M. Simmie, Combust. Flame 124

(1-2) (2001) 326-329.

9. V. P. Zhukov, V.A. Sechenov and A.Y. Starikovskii, Combustion Explosion

and Shock Waves 39 (5) (2003) 487-495.

10. G. P. Smith, D. M. Golden, M. Frenklach, N. W. Moriarty, B. Eiteneer, M.

Goldenberg, C. T. Bowman, R. K. Hanson, S. Song, W. C. Gardiner, V.

Lissanski and Z. Qin, http://www.me.berkeley.edu/gri_mech/..

11. A. Lifshitz, K. Scheller, A. Burcat and G. B. Skinner, Combust. Flame 16 (3)

(1971) 311-321

12. R. K. Cheng and A. K. Oppenheim, Combust. Flame 58 (2) (1984) 125-139.

13. J. Huang, W. K. Bushe, P. G. Hill and S. R. Munshi, Int. J. Chem. Kin. 38 (4)

(2006) 221-233.

14. J. Huang, P. G. Hill, W. K. Bushe and S. R. Munshi, Combust. Flame 136

(1-2) (2004) 25-42.

15. C. T. Bowman, R. K. Hanson, D. F. Davidson, W. C. Gardiner, V. Lissanski,

G. P. Smith, D. M. Golden, M. Frenklach and M. Goldenberg,

http://www.me.berkeley.edu/gri_mech/.

16. K. J. Hughes, T. Turanyi, A. R. Clague and M. J. Pilling, Int. J. Chem. Kin.

33 (9) (2001) 513-538.

Page 76: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 3

71

17. O. Conaire, H. J. Curran, J. M. Simmie, W. J. Pitz, C. K. Westbrook, Int. J.

Chem. Kin. 36 (11) (2004) 603-622.

18. J. Li, Z. W. Zhao, A. Kazakov and F. L. Dryer, Int. J. Chem. Kin. 36 (10)

(2004) 566-575.

19. E. L. Petersen, D. M. Kalitan, S. Simmons, G. Bourgue, H. J. Curran and J.

M. Simmie, Proc. Combust . Inst .31 (2007) 447-454.

20. D. Lee and S. Hochgreb, Combust. Flame 114 (3-4) (1998) 531-545.

21. S. Tanaka, F. Ayala and J. C. Keck, Combust. Flame 133 (4) (2003) 467-

481.

22. J. Strohle and T. Myhrvold, Int. J. hydrogen energy 32 (2007) 125-135.

Page 77: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

72

Chapter 4 One-dimensional laminar flames Experimental Techniques, Procedures and Burner Setup

Page 78: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

73

4.1 General introduction

In this chapter, the experimental procedures for the measurement of temperature

and concentrations of HCN and C2H2 in one-dimensional flames are described.

The flame temperature is measured using Coherent Anti-Stokes Raman

Scattering (CARS) for N2 thermometry. This in-situ technique has been demonstrated

to provide an accuracy of few tens degrees with high spatial and temporal resolution,

see for example [1,2], and is one of the best methods for flame thermometry [3]. The

experimental setup and measurement procedure was essentially identical to that

reported elsewhere [4].

In contrast to the “standard” method of CARS for temperature measurements,

the low concentrations of HCN and C2H2 in flames are more difficult to measure.

Various methods have been developed to detect HCN and C2H2 in combustion

systems, among them intrusive techniques (extractive probe sampling) and in-situ

laser diagnostic techniques. Extractive probe sampling and subsequent analysis of the

sample, using infrared spectroscopy or mass spectrometry, offer the advantage of

flexibility and relatively simple and accurate methods to analyze the flame samples.

One of the drawbacks of extractive probe sampling in flames is the potential distortion

of species profiles [5]. This problem can be avoided by the use of in-situ laser

diagnostic approaches such as laser-induced fluorescence (LIF), spontaneous Raman

scattering, direct absorption spectroscopy and many others. Each of these techniques

has its strengths and weaknesses. Spontaneous Raman scattering and LIF have both

been successful at detecting C2H2 in flames without significant interference from

other flame species [6,7]. However, the current state of these techniques is such that

they are not yet generally applicable for measurements under flame conditions. The

sensitivity of the Raman technique is low [6] and therefore only applicable in

environments of high C2H2 concentration. Measurement of C2H2 using LIF also

suffers from low signal levels and requires further research to quantify the measured

signal [7]. Although acetylene absorbs strongly in the infrared (IR) spectral range,

selective detection using a direct absorption technique is difficult, due to interference

from other flame molecules such as CO2, CO, and H2O. As an illustration of possible

interferences in the IR spectral range, figure 4.1a shows a part of the calculated

absorption spectrum of a gas mixture of acetylene and water at typical flame

conditions, using data from refs. [8-11]. The figure shows that the acetylene

Page 79: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

74

absorption lines are impossible to resolve due to the large number of strong and

overlapping water lines.

Figure 4.1. Calculated absorption spectra for a gas mixture of 0.3% C2H2 and 18% H2O in N2 at 1800K (a) and 300K (b). Line positions, strengths and broadening coefficients are taken from [8-10] and [11], respectively. The in-situ detection of low concentrations of HCN in flames is difficult as well,

since it only absorbs strongly in the infrared spectral range. To our knowledge, the

only in-situ technique used to measure HCN in a low pressure flame is fiber laser

Page 80: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

75

intracavity absorption spectroscopy (FLICAS) [12], applied in the IR-region around

1.5 μm. The flame study was doped with NH3, which increased the HCN

concentration in the flame substantially. The concentration of native HCN in

hydrocarbon flames is much lower, and, similar to acetylene, the absorption lines of

HCN in these flames will be difficult to distinguish from other flame molecules.

Consequently, the concentrations HCN and C2H2 in the flames we are interested in are

too low to detect accurately using in-situ absorption techniques. At room temperature,

where the HCN and C2H2 absorption features are much stronger and H2O and other

interfering lines are less intense, this results in several HCN and C2H2 absorption

peaks with negligible interference. This is illustrated for the gas mixture of C2H2

(0.3%) and H2O (18%) at room temperature in figure 4.1b. To take advantage of this

situation, we thus developed an extractive probe sampling system with room

temperature analyses, using tunable diode laser absorption spectroscopy (TDLAS) at

∼1.5 μm to measure concentrations of HCN and C2H2. We analyze probe

perturbations to minimize and correct for distortion of the species profile arising from

the presence of the probe sampling system in the flame.

Page 81: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

76

4.2 Burner

The measurements described in this thesis were performed in one-dimensional,

atmospheric-pressure premixed flames stabilized above a McKenna products burner

of 60 mm diameter, shown schematically in figure 4.2. The burner consists of a

sintered bronze plug, which serves as a flame holder, and is surrounded by another

sintered section for a shroud gas. In this study a nitrogen shroud was used to prevent

air entrainment in the combustion products. A cylindrical chimney with a 60 mm

inner diameter was positioned approximately 30 mm above the burner to stabilize the

column of post-flame gases. The burner surface is cooled by water flowing through

coils imbedded in the sintered flame holder.

Figure 4.2. Schematic of the McKenna Products burner.

The burner was affixed to a positioner that moves the burner with a precision of

0.1 mm, to allow measurement of axial profiles of species concentration and

temperature.

Page 82: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

77

4.3 Gas handling system

Figure 4.3. Experimental set-up used for transport, mixing, and analysis of the fuel-air mixture.

Figure 4.3 shows the flow scheme for the burner feed gases and the system for

analyzing the composition of the unburned fuel-air mixture. The flow rates of all

gases were measured by calibrated mass-flow meters (Bronkhorst, EL-FLOW),

digitized by an analog-digital converter and processed by a PC. The flow ranges of the

meters were selected to provide an accuracy of better than 5%. All fuels used in this

study were supplied in cylinders with purity better than 99.99% and dry, filtered air

was supplied by an oil-free compressor. The gases were mixed in a tube 50 cm long,

and a small part was diverted to a Maihak Unior 610 gas analyzer for measuring the

methane and oxygen mole fractions in the fuel-air mixture, while the rest is supplied

to the burner system. The by-pass burner makes it possible to vary the flow rate

through the burner during the experiment without changing the choked flows of the

individual gases.

The equivalence ratio ϕ of the fuel-oxidizer mixtures was calculated according

the following expression (described in detail in chapter 1),

Page 83: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

78

24 ,2

2

,2

4 1.][][1.

][][

HstCHst fOH

fOCH

+=ϕ . (4.1)

Whereas it is possible to determine the equivalence ratio through measuring the flow

rates, to increase accuracy we determined ϕ based on measuring CH4 and O2 mole

fractions in the unburned mixture using the Maihak analyzer. In CH4/air flames,

measuring only the CH4 mole fraction in the cold mixture is sufficient to calculate the

complete mixture composition, since the oxygen/nitrogen ratio in the mixture known.

This procedure provided accuracy better than 2% for all equivalence ratios used. To

determine the composition of the cold unburned mixtures of the CH4/H2/air flames,

the CH4 and O2 mole fractions are both measured using the Maihak gas analyzer, the

N2 mole fraction is calculated from the measured O2 mole fraction and the known

ratio [O2]/[N2] in air, and the H2 mole fraction is calculated from the balance,

air

measmeasmeas ONOOCHH ⎟⎟

⎞⎜⎜⎝

⎛−−−=

][][*][][][1][

2

2.2.242 . (4.2)

Based on this method, the accuracy of the equivalence ratios determined was better

than 5% in all CH4/H2/air flames measured. Detailed information on the uncertainty

analyses can be found in [13]. The mass flux is calculated from the measured mass

flow, assuming the surface area of the burner is equal to that of the flame. The

accuracy of the mass flux through the burner surface was estimated approximately

10%, determined by the uncertainties in the measured mass flow rate and the

uncertainty in the flame area.

Page 84: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

79

4.4 Extractive-probe sampling system

Figure 4.4. Schematic of the microprobe sampling system.

Figure 4.4 shows a schematic of the probe sampling system. All flames were

sampled by a quartz microprobe having a similar design to that described in [14], with

an orifice diameter of about 100 μm. For the acetylene measurements, with exception

of the probe tip, the quartz probe was cooled over a distance of 35 cm by water at

12οC. After passing an ice-cooled water trap, the sampled gas flowed through a 1 m-

length stainless-steel tube sealed on both ends by quartz windows to form a

measurement cell. The pressure in the tube was monitored by an electronic pressure

transducer and kept constant at 100 mbar by a vacuum pump installed in the exit of

the sampling system; this provided rapid removal and quenching of the gas sample.

Because HCN is soluble in water, measurement of this component was performed

using an uncooled probe and water at room temperature in the cooling trap. In this

case no condensation of water from the combustion gases in the cooling trap was

observed during sampling.

Page 85: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

80

4.5 Estimate of the conversion of C2H2 and HCN during sampling

To estimate the degree of conversion of C2H2 and HCN to other products during

sampling, the probe is modeled as a closed system containing a homogeneous gas

mixture, described in detail in chapter 1. The model uses a prescribed cooling rate,

based on characteristic time scales in the experimental set-up to calculate the mixture

composition history profile. The probe modeling was divided in two parts: rapid

adiabatic expansion in the probe tip (τ1) followed by cooling of the sample at constant

pressure (τ2). During expansion in the probe tip the gas sample undergoes a rapid

decrease in pressure and temperature, described by,

1/exp)( .0.max

τtPPP

−−= . (4.3)

The temperature drop associated with the pressure drop is calculated in the program

using equation (2.2), by assuming isentropic flow conditions. The characteristic

cooling time in the probe tip used was τ1 =10-4 s, which is one order longer than

estimated by assuming critical flow in the probe orifice. The initial composition and

temperature were varied in the model using the results of one-dimensional flame

calculations, described in detail in chapter 1. The temperature (T0) of the gas flow in

the probe immediately after expansion is taken from the temperature calculations in

the probe tip at a probe backpressure (P0) of 0.1 bar and put in the model. Further

downstream of the probe tip the gas is cooled at constant pressure (P0=0.1 bar) to

room temperature (Tmin) by heat transfer to the coolant, according to,

.min/

min0 2exp)( TTTT t +−= − τ (4.4)

In the model we used τ2=0.5s; this value is one order higher than that based on heat

transfer estimations. GRI-Mech 3.0 [15] is used as chemical mechanism in the

calculations.

Page 86: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

81

Figure 4.5. a) Calculated temperature and concentration curves during expansion in the probe tip, using τ1=10-4s model. The initial temperature and composition are taken from a simulated adiabatic (free) premixed methane/air flame with φ=1.4 and mass flux 0.014g/cm2s, at 3mm from the burner surface. b) Calculated temperature and concentration curves downstream the probe tip (P=0.1 bar), using τ2=0.5s.

As an example, figure 4.5a presents the predicted C2H2 conversion during

expansion in the probe tip and 4.5b shows the predicted C2H2 conversion further

downstream the probe tip during sampling, for the case in which the conversion was

maximal. The calculations show that in the present experimental setup, for all flames

under investigation, the conversion of C2H2 in the probe is less than 15%, and less

than 10% for HCN, at axial distances greater than 2.5 mm from the burner surface.

We shall return to complications of sampling in Chapter 5.

Page 87: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

82

4.6 Laser absorption spectroscopy

4.6.1 Theory If laser radiation of intensity I0(v) at frequency v passes through an absorbing

medium over a length l, the intensity of the transmitted laser beam I(v) can be

expresses according to the Lambert-Beer law:

lvvIvI

)()()(

ln 0 α=⎟⎟⎠

⎞⎜⎜⎝

⎛, (4.5)

where α(v) is the absorption coefficient. The absorption coefficient is proportional to

the concentration of the absorbing molecules and is dependent upon the temperature;

it can written as [16]

mPxvfTSv )()()( =α , (4.6) where S(T) is the line intensity or integrated absorption coefficient per unit pressure

(expressed in cm-2atm-1), f(v) the spectral line function (normalized such that

∫∞

∞−

= 1)( dvvf ), P the pressure and xm the species mole fraction. The line function f(v)

depends on temperature and the collisional environment of the molecule. The shape of

f(v) depends upon the type of broadening. Doppler broadening is due to thermal

motion of molecules and collisional broadening is a result of perturbations in the

energy levels of absorbing molecules caused by collision with the other molecules.

Since the parameters of the line function f(v) are often unknown, it is better to

integrate the left and right parts of equation (4.5) over the entire line profile and

substitute (4.6) into (4.5) in order to determine the species mole fraction:

PlTS

dvvIvI

xm )(

)()(

ln 0∫∞

∞−= . (4.7)

Page 88: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

83

The smallest detectable absorption is limited by the minimum difference in I and I0

that can be measured, being determined by noise in the measurement system. Typical

absorption sensitivities using direct absorption allow the detection of absorbances

(0

lnII ) of the order of 10-3. For smaller absorbance it is difficult to reach an adequate

signal-to-noise ratio to distinguish between Io and I.

Whereas the mole fraction of acetylene is expected to be large enough to be

measured by direct absorption (using TDLAS, see below), we anticipate that this will

not be the case for HCN. One-dimensional flame calculations (chapter 1) using GRI

Mech 3.0 [15] indicate ppm level mole fractions of HCN. For such mole fractions the

calculated absorbance, using line positions, strengths and broadening coefficients

from [17] and [10] respectively is on the order of 10-4. It is clear that the detection

sensitivity should be increased to resolve low concentrations of HCN with reasonable

accuracy. To increase the sensitivity, several detection techniques have been

developed over the past decades. These include signal enhancement methods such as

multipass cells [18,19] and cavity-enhanced spectroscopy [20], and noise-reduction

methods like wavelength- and frequency-modulation absorption spectroscopy [21]. In

this study, modulation spectroscopy with second harmonic detection is implemented

to increase the detection sensitivity. Wavelength-modulation absorption spectroscopy

(WMAS) is well described in the literature [21,22] and is only briefly described

below.

4.6.2 Wavelength Modulation Absorption Spectroscopy (WMAS)

The measured intensities always contain a certain amount of noise, which limits

the smallest detectable absorption. The dominating source of noise is often 1/f noise

[23], which is larger at low frequencies and will decrease at higher frequencies.

Wavelength-modulation absorption spectroscopy (WMAS) is an effective technique

to reduce the noise, and thereby increase the detectability. The principle of WMAS is

smooth modulation of the wavelength at a certain frequency followed by detection of

the modulated laser radiation at the original modulation frequency or at one of its

harmonics. One of the advantages of modulation spectroscopy is that only the noise

centered on the detection frequency will influence the measurements. The detection is

Page 89: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

84

shifted to higher frequencies where the 1/f noise is smaller. The basic approach is to

modulate the wavelength of the laser radiation around its center (here using the

frequency of the radiation, vc, in the derivation) at frequency fm with a modulation

amplitude va. The instantaneous frequency v(t) of the laser radiation can than be

represented as,

tfvvtv mac π2sin)( += . (4.8)

After absorption in the medium, the intensity of the modulated, transmitted laser

beam I(v(t)) and the reference laser beam I0(v(t)) are related by the Lambert-Beer law,

already introduced in (4.5). WMAS is usually used for samples that have low

absorptions. This allows expanding the exponent in equation. (4.5) in a Taylor series.

By neglecting the higher order terms in the Taylor series one obtains:

]))((1))[(())(( 0 LtvtvItvI α−= . (4.9)

The instantaneous laser intensity I(vc+vasin2πfmt) can be expanded in a Fourier sine

series;

∑∞

=

=+0

2sin),()2sin(n

macnmac tnfvvHtfvvI ππ . (4.10)

The nth harmonic component of the intensity of the transmitted laser beam Hn(vc,va),

for n≥1 is expressed as

1( , ) ( sin 2 )sin 22

t

n c a c a m mt

H v v I v v f t f ntdtπ ππ

= +∫ . (4.11)

To simplify the analyses we assume that the laser intensity is independent of the

frequency. In this case Hn can be written as,

Page 90: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

85

0 ( sin 2 )sin 22

t

n a m mt

LIH v v f t f ntdtα π ππ

= − +∫ . (4.12)

It can be seen from equation (4.12) that Hn is directly proportional to the

absorption coefficient α(vc+vasin2πfmt) and thus each harmonic component is directly

proportional to the species concentration in the absorption layer. Figure 4.6 shows a

schematic illustration of a broadened absorption line and the calculated corresponding

first (n=1) and second (n=2) harmonics (Fourier components).

Figure 4.6 Principle of WMAS with harmonic detection.

Any of the harmonics (n=1,2…) can be used; however optimum signals are

generated using second-harmonic detection [21]. The increased sensitivity of the

second-harmonic detection technique arises in part from the elimination of featureless

background signals. For detecting weak absorptions it is important to maximize the

harmonic signals by optimizing the modulation amplitudes. The simulated second

harmonic signals shown in figure 4.7 are calculated for different modulation

Page 91: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

86

indices2/1v

vm a

Δ= , using an exact solution to the integral in equation (4.11) for a

Lorentzian line shape [24]. Here, Δν1/2 is the half width at half maximum (HWHM) of

the absorption line.

Figure 4.7 Typical 2f-lineshape for different modulation indices.

As can be seen from figure 4.7, the 2f-line shapes becomes broader with

increasing modulation index and reaches the maximum peak height at a modulation

index m∼2. The maximum peak height occurs at m∼2 for all line shapes [24,25]. In

environments where the absorption is very low, the modulation index is often set to a

value close to m∼2; for cases in which it is necessary to reduce interferences from

nearby transitions, the modulation index is decreased to minimize the broadening of

the 2f-line shape.

Page 92: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

87

4.7 Experimental setup for Tunable Diode Laser Absorption Spectroscopy

Figure 4.8. General schematic of the experimental TDLAS set-up.

Figure 4.8 shows the basic schematic of the experimental setup for Tunable

Diode Laser Absorption Spectroscopy (TDLAS) used for direct absorption

measurements of acetylene. The radiation from a New Focus 6326 tunable diode

laser, with a linewidth less than 300 kHz, was directed through the absorption cell.

Before entering the cell, a part of the laser radiation was split off to produce the

reference signal. The powers of the reference and sample beams were measured by

New Focus 2033 large area photodiodes with internal amplifiers. The photodiode

signals were digitized and processed by a PC. The laser wavelength was swept over

the span of 30 GHz with a scan rate 150 MHz/s by applying a voltage to the

piezoceramic plate, which moves the tuning end-mirror inside the laser. The

measurements were performed in the region around 1530 nm, where the P (9)

absorption line of the ν1 + ν3 band of C2H2 is located [8]. This line was selected

because of its relatively high oscillator strength [9] and lack of interference from

transitions of other flame molecules.

Page 93: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

88

4.7.1 Experimental procedure TDAS measurements of acetylene Whereas it is possible to derive absolute values of acetylene concentration from

the measured integral absorption coefficient (equation 4.7), direct calibration by a gas

with known acetylene concentration was used in present work to enhance the

accuracy. This method avoids the uncertainties arising from the non-infinite limits of

integration when deriving the integrated absorption coefficient, and from converting

the voltage applied to the piezoceramic plate to the wavelength shift. The calibration

procedure is performed by measuring the absorption coefficient of a known amount of

acetylene under the same experimental conditions as those existing for the sampled

gas. As a typical example, Figure 4.9 shows the dependence of the logarithm of the

ratio of the reference to transmitted signal (absorbance) upon the laser wavelength

(expressed as the applied voltage) when the measurements were performed in a

mixture containing 5000 ppm C2H2 in N2 at 0.1 bar.

Figure 4.9 experimental absorption profile obtained in a mixture containing 5000 ppm C2H2 in N2 in the vicinity of the P(9) line of ν1+ν3 band.

When comparing the measurements performed in the calibration gases with those of

the combustion products, no differences were found in absorption line shapes. This

implies that the spectral line function in equation (4.6) is independent of the gas

composition in the current experiments. This result is perhaps to be expected, since

Page 94: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

89

the broadening efficiencies of C2H2 and N2 for the acetylene lines are comparable

[10], while the concentrations of the other gases with unknown broadening efficiency

are low.

Taking into account that the spectral line function is the same in the calibration

gas and in combustion products one derives from equation. (4.7),

⎟⎟⎠

⎞⎜⎜⎝

⎟⎟⎠

⎞⎜⎜⎝

=

⎟⎟⎠

⎞⎜⎜⎝

⎟⎟⎠

⎞⎜⎜⎝

=

i ical

ical

i i

i

calv

v cal

lcal

v

vcalM

vIvI

vIvI

x

dvI

vI

dvvIvI

xx

)()(

ln

)()(

ln

)(ln

)()(

ln

0

0

.2

1

.0

2

1

0

. , (4.13)

where the subscripts cal and i refer to the calibration gas and i-th measured point,

respectively. In the current experimental setup, for acetylene mole fractions above

1000 ppm, the accuracy of the measured C2H2 concentrations is ~5%, and is

determined mainly by the uncertainty in the calibration gas concentration. At low

acetylene mole fractions, the uncertainties in the measured integral absorption

coefficient become dominant, which results in deteriorating accuracy, up to 15% at

100 ppm C2H2. While equation (4.7) only requires one calibration point to determine

the measured acetylene concentration, to verify the possible influence of non-linearity

in the detection system, the integral absorption coefficient was measured in

calibration gases with different C2H2 concentrations. As can be seen from figure 4.10,

showing the results of these measurements, the detection system possesses excellent

linearity: deviation of the fitted and experimental values is less than 5%.

Page 95: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

90

Figure 4.10. Integrated absorption coefficient of the C2H2 P(9) line of the ν1+ν3 band as a function of acetylene mole fraction diluted in N2. Solid line is a linear fit.

The measured C2H2 mole fractions were recalculated to flame conditions by

taking into account the partial removal of water in the ice-cooled trap. The residual

water concentrations in the absorption tube were determined from the integrated

absorption coefficient of the water absorption line at 6755.02 cm-1 [11] using equation

(4.7). The line intensity S(T) of the water line was taken from [11].

4.8 Experimental procedure WMAS with second harmonic detection

4.8.1 HCN measurements The experimental sep-up used for the WMAS experiments with second harmonic

detection is almost identical to that presented schematically in figure 4.8. The only

difference is that we do not use the reference beam. The laser frequency v is

modulated around the center frequency vc by applying a sinusoidal voltage from an

external generator to the piezoceramic plate which moves the tuning end mirror inside

the laser. The modulation depth va, and frequency fm, in equation (4.8) are chosen to

be ∼0.67 GHz and 0.5 kHz, respectively, to provide maximal sensitivity for the

present experimental setup [21]. The power of the transmitted laser beam is digitized

Page 96: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

91

by an Agilent 54830B oscilloscope with sampling rate fs = 50 kHz and processed by a

PC. The second harmonic amplitude I2f(vc) is calculated from the sampled signal

series using the following expression:

21

21( ) ( ) sin(2 )N

mf c i c

i s

fI v I v iN f

π=

= ⋅∑ . (4.14)

The summation (4.14) is performed with the number of samples N ~ 10000, which

corresponds 0.2 s total sampling time. Further increasing the number of samples does

not result in substantial improvement in the signal-to-noise ratio. The HCN

measurements were performed in the region around 1545 nm where the P(13)

absorption line of the ν2 band of HCN is located [17]. This line was selected because

of its relatively high oscillator strength and lack of interference from other flame

molecules. To cover the entire absorption line, the center frequency vc is tuned in

steps of 0.12 GHz over span of ∼10 GHz by changing the mean voltage applied to the

piezotransducer. The improvement in signal-to-noise ratio obtained using WMAS as

compared to direct absorption is illustrated in figure 4.11. The measured direct

absorption profile measured in a gas mixture containing 90 ppm HCN in N2 is shown

in figure. 4.11a. Figure 4.11b shows a typical HCN second harmonic spectrum in the

same gas mixture. As can be seen, the signal-to-noise ratio is improved by more than

one order by using WMAS with second harmonic detection.

Page 97: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

92

Figure. 4.11a.) Direct absorption scan of 90 ppm HCN in nitrogen in the vicinity of the P(13) line of ν2 Σ-Σ band, b.) Second harmonic signal in the same mixture for the same absorption transition.

Although it is possible to derive absolute HCN concentrations by fitting the

measured second harmonic absorption profiles, we use signal calibration in gases with

known HCN mole fractions, as done for acetylene above. In this approach, knowledge

of the absorption line shape is not required, which substantially improves the accuracy

Page 98: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

93

of measured HCN concentrations. Below, we give a short description of the

calibration procedure, starting by presenting the absorption coefficient α(v) as the sum

of the absorption coefficient αM of the HCN molecules and background absorption

αBG.

)()()()()()( vPxvgTSvvv BGmBGml αααα +=+= . (4.15)

Substituting (4.15) into the Lambert-Beer law for optically thin absorption layers (4.9)

and taking the Fourier transform, while taking into account that according (4.8) v is a

function of time, we receive the following expression for the Fourier transform of the

transmitted signal at frequency f,

0 0 0( ) ( ) ( ) ( ) ( ) MBGI f I f I f l I g f S T Px lα= − −% , (4.16)

where the tilde above the variable denotes its Fourier transform. As we can see from

this expression, the measured signal consists in general of three terms. The first two

terms in (4.16) are due to the frequency dependence of the laser power and

background absorption coefficient. In the case of sinusoidal modulation and second

harmonic detection (f = 2fm), these terms will vanish only when laser power and

background absorption coefficient are constant or one of them is linearly dependent

upon the laser frequency v. The term 0 0(2 ) (2 )m mBGI f I f lα− can be determined by

measuring the second harmonic signal in media without HCN molecules. The

measurements, performed at a height of 10 mm above the burner surface in a

methane-air flame with equivalence ratio ϕ = 1.3 and mass flux ρν=0.015g/cm2s

where HCN concentration is expected to be very low, showed that in the spectral

region used in the present work the background 0 (2 )mBGI fα does not exceed the

signal from HCN at mole fraction of 5 ppm. Because the concentrations of main flame

components that contribute to the background absorption coefficient (H2O, CO and

CO2) vary only slightly with equivalence ratio, the value of 0 0(2 ) (2 )m mBGI f I f lα−

was used for deriving HCN concentrations in all CH4/air flames studied here. To

determine the background for the CH4/H2/air flames, a similar reference flame (ϕ=1.3

Page 99: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

94

and ρν=0.015g/cm2.s) was used with the same percentage of hydrogen added as for

the flames under investigation.

A typical second harmonic spectrum of HCN, measured in a flame at ϕ=1.4, at 2

mm above the burner surface, is presented in figure 4.12a. The second harmonic

spectrum of HCN ( 0 (2 ) ( )m mI g f S T Px l ), shown in figure 4.12c, was obtained by

subtracting the measured second harmonic signal of the reference flame

( 0 0(2 ) (2 )m mBGI f I f lα− ), given in figure 4.12b, from that shown in figure 4.12a. In

figure 4.12c, we can see the characteristic shape expected for the second derivative of

the bell-shaped spectral line profile.

Using the second harmonic signal to derive quantitative information on species

concentrations is rather difficult. Direct fitting according to the expression (4.16)

requires knowledge of functional dependence of both absorption spectral line and

laser intensity upon frequency. Moreover, direct integration of the second harmonic

signal would yield a value close to zero. To overcome these difficulties, the extracted

second harmonic signal is integrated over a sufficiently large region [v1, v2] in the

vicinity of the HCN spectral line. From the expression (4.16) it follows that

2

0 0

1

(2 ) (2 ) (2 )v

m m m mBGv

x C I f I f I f l dvα= − −∫ % , (4.17)

where the coefficient C is expressed as

2

0 )

1

1

( ) (2v

mv

C

S T Pl I g f dv

=

∫. (4.18)

It should be pointed out that, in general, the coefficient C depends upon

composition of the sampled gas. When comparing the measurements performed in the

calibration gases with those from the combustion products, no differences were found

in the second harmonic line-shapes. This implies that the spectral line function is

independent of the gas composition in the present experiments, and therefore the

coefficient C can be regarded as constant.

Page 100: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

95

Figure 4.12a) 2f-signal measured in a flame with ϕ=1.4 and ρν=0.01g/cm2.s at 2mm from the burner surface, b) 2f signal measured in the reference flame (ϕ=1.3, ρν=0.15g/cm2.s at 10mm from the burner surface), c) Extracted second harmonic spectrum of HCN.

Page 101: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

96

Equation (4.17) was verified by measurements of the second harmonic signals at

different HCN concentrations in N2. The results of these measurements, presented in

figure 4.13, show a linear dependence between the integrals of the absolute value of

( 0(2 ) (2 )m mI f I f−% and the HCN mole fraction. Clearly visible is the offset at zero

HCN concentration; this is mainly due to the noise in the measured signals. When the

second harmonic signal is close to zero, the integration is performed over absolute

values of the noise, which results in a non-zero integral. In the current experimental

setup, when HCN mole fractions are above ~10 ppm, the accuracy (~10%) of the

measured HCN concentrations is determined mainly by the uncertainty in the

calibration gas concentration. At lower mole fractions the uncertainties in the

measured integral absorption coefficient become dominant, which results in

deteriorating accuracy, up to 30% at 3 ppm HCN.

Figure 4.13. Integrated absolute value of the second harmonic signal in the vicinity of the P(13) line of ν2 Σ-Σ band as a function of HCN concentrations in N2.

Page 102: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

97

Literature

1. A. V. Mokhov and H. B. Levinsky, Proc. Combust. Inst. 26 (1996) 2147-

2154.

2. A. C. Eckbreth, Laser Diagnostic for Combustion Temperature and Species,

(2nd Edition, Gordon and Breach, Amsterdam, 1996).

3. "Applied Combustion Diagnostics," K. KohseHoinghaus and J. B. Jeffries,

eds., (Taylor & Francis, New York, 2002).

4. V. V. Toro, Experimental study of the structure of laminar axisymetric H2/air

diffusion flames, Ph.D. Thesis, RUG, 2006 (ISBN 90-367-2703-0).

5. E. L. Knuth, Combust. Flame 103 (3) (1995) 171-180.

6. A. V. Mokhov, S. Gersen, H. B. Levinsky, J. Chem. Phys. Let., 403 (2005)

233-237.

7. B. A. Williams., J. W. Fleming., Appl. Phys .B. 75 (2002) 883-980.

8. Q. Kou, G. Guelachvili, M. A. Temsamani, M. Herman, Can. J. Phys. 72

(11-12) (1994) 1241-1250.

9. R. El Hachtouki, J. Vander Auwera, J. Mol. Spectrosc. 216 (2) (2002) 355-

362.

10. A. S. Pine, J. Quant. Spectrosc. Radiat. Transfer 50 (2) (1993) 149-166.

11. R. A. Toth., Appl.Opt., 33 (1994) 4852.

12. A. Goldman., I.Rahinov., S. Cheskis., B. Lohden., S. Wexler., K. Sengstock.,

V. M. Baev, Chem.Phys.Let., 423 (2006) 147-151.

13. A. Sepman, Effect of burner stabilization on nitric oxide formation and

destruction in atmospheric pressure flat premixed flames, Ph.D. Thesis,

RUG, 2006 (ISBN 90-367-2702-2).

14. R. M. Fristrom and A. A. Westenberg, Flame Structure, (McGraw-Hill, New

York, 1965).

15. G. P. Smith, D. M. Golden, M. Frenklach, N. W. Moriarty, B. Eiteneer, M.

Goldenberg, C. T. Bowman, R. Hanson, S. Song, W. C. Gardiner, V.

Lissanski, Z. Qin, http://www.me.berkeley.edu/gri_mech/.

16. A. C. Eckbredt, Laser diagnostic for combustion temperature and species,

(2nd Edition, Gordon and Breach, Cambridge, 1996).

17. A. M. Smith, S.L. Coy, W. Klemperer, K.K. Lehmann, J. Mol. Spectros.

134 (1) (1989) 134-153.

Page 103: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 4

98

18. R. G. Pilston, J. U. White: A long path gas absorption cell, J. Opt. Soc. Am.

44, 572–573 (1954).

19. J. U. White: Long optical paths of large aperture, J. Opt. Soc. Am. 32, 285–

288 (1942).

20. A. O’Keefe, D. A. G. Deacon: Cavity ring-down optical spectrometer for

absorption measurements using pulsed laser sources, Rev. Sci. Instrum. 59,

2544 (1988).

21. J.A. Silver, Applied Optics, 31 (6) (1992) 707-717.

22. P. Kluczynski, J. Gustafsson, A. M. Lindberg, O. Axner, Spectrochimica

Acte Part B, 56 (2001) 1277-1354.

23. C. Th. J. Alkemade, T. J. Hollander, W. Snelleman, P. J. Th. Zeegers,

International Series in Natural Philosophy 3 (1982) 272 .

24. R. Arndt, Appl. Opt. 36 (8) (1965) 2522-2524.

25. J. Reid, D. Labie, Appl. Phys. B 26 (1981) 203-210.

Page 104: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 5

99

Chapter 5

Extractive Probe Measurements of Acetylene in Atmospheric-Pressure Fuel-Rich Premixed Methane/Air Flames

Page 105: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 5

100

5.1 Introduction

As discussed in previous chapters, one of the major advances of the last decades

in the combustion science is the prediction of flame structure by numerical

simulations using detailed transport and chemical mechanisms. Because of the

complexity of these mechanisms and uncertainties in the rates of the key chemical

reactions, the predictive power of the numerical simulations can be tested only by

comparing calculated and measured flame properties under well-defined experimental

conditions. The comparison of the spatial profiles of intermediate species is

particularly important for testing the adequacy of chemical mechanisms. One of the

key intermediates in many high temperature processes is acetylene (C2H2), which

plays important role in the formation of polycyclic aromatic hydrocarbons and soot in

hydrocarbon combustion [1-3] and in the chemical vapor deposition of diamond [4].

Acetylene has been extensively investigated in both atmospheric- and low-pressure

flat premixed flames [5-11]. At atmospheric pressure, large discrepancies have been

observed [8,11] between measured results and those calculated based on the C2H2

submechanism derived from Miller and Mellius [12]. However, the acetylene

measurements in these studies were performed using extractive probe sampling,

which as discussed in Chapter 4 has a serious drawback, i.e., the distortion of the

composition and temperature profiles in the flame. Estimating the magnitude of this

distortion, for example, from chemical reactions on the probe surface or acceleration

of the combustion products into the probe orifice are rather difficult [13]. Moreover,

these estimates (as was done in Chapter 4) require detailed knowledge of the kinetics

of the chemical reactions involving the measured species that itself is the subject of

investigation. These complications necessitate the verification of the results obtained

by the extractive probe by an independent technique. Recently, we have reported the

measurement of native C2H2 in a fuel-rich methane/air flame at equivalence ratio

ϕ = 1.55, using spontaneous Raman scattering [14]. This method thus provides us

with the means to verify the results of extractive probe sampling for acetylene

measurement, and to deliver reliable experimental results regarding C2H2 formation

and destruction in atmospheric-pressure methane/air flames.

Towards this end, we have measured the profiles of C2H2 mole fraction in flat

atmospheric-pressure rich-premixed methane/air flames using both spontaneous

Raman scattering and microprobe gas sampling followed by tunable diode laser

Page 106: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 5

101

absorption spectroscopy (TDLAS). These measurements are supplemented with

profiles of flame temperature, obtained using coherent anti-Stokes Raman Scattering

(CARS), and the experimental results are compared with the predictions of one-

dimensional flame calculations.

5.2 Experimental

Here we briefly summarize the experimental method; Chapter 4 discusses the

methods in more detail. The measurements were performed in atmospheric-pressure

methane/air flames stabilized above a McKenna Products sintered bronze burner of

60 mm diameter. To prevent air entrainment in the combustion products a nitrogen

shroud was used. The flame was stabilized by a cylindrical chimney with a 60 mm

inner diameter, which was positioned approximately 30 mm above the burner surface.

The flame temperature was varied by changing the mass flow through the burner and

measured by broadband planar BOXCARS for nitrogen thermometry. Details of the

CARS experiment are described elsewhere [15]. The flow rates of methane and air

were measured by calibrated mass flow meters and the equivalence ratio was

determined by measuring the methane concentration in the unburned fuel-air mixture.

For calibration purposes, nitrogen doped with a known amount of acetylene was

flowed through the burner instead of the methane-air mixture. Measurements were

obtained at different axial positions in the flame by moving the burner with a

precision positioner relative to the laser beams and sampling probe in steps of 1 mm.

The flames were sampled by a cooled quartz micro-probe, and the sampled gas

flowed through an absorption cell and analyzed using TDLAS. As discussed in

Chapter 4, estimates of C2H2 conversion during sampling indicated that in the present

experimental setup conversion of acetylene in the probe is less than 15% when

sampling is made at axial distances greater than 2.5 mm from the burner surface.

These estimates are supported by measurements at different suction backpressures,

which showed no significant changing in the measured HCN concentration when

varying pressure from 0.05 to 0.35 Bar.

Page 107: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 5

102

5.3 Results and discussion

The measurements were performed in a set of fuel-rich flames with different

equivalence ratios and mass fluxes. The flame parameters (equivalence ratios, mass

fluxes and temperatures at 5 mm above the burner surface) are presented in Table 5.1.

Table. 5.1. Flame parameters

Flame ϕ ρv, g/cm2·s) T, K

A 1.5 0.005 1763

B 1.5 0.007 1835

C 1.5 0.008 1852

D 1.45 0.007 1833

E 1.45 0.0085 1885

F 1.45 0.010 1916

G 1.4 0.005 1762

H 1.4 0.007 1816

I 1.4 0.0085 1850

The temperature measurements showed that all the flames studied had a domain with

constant temperature extending at least 20 mm radially from the centerline, and from

3 mm to 15 mm in the axial direction. As typical examples of the temperature

measurements, the radial profile at height 10 mm above the burner surface and

centerline axial profile in flame A are presented in Figs. 5.1 and 5.2, respectively. The

radial profile shows a core region of ~ 20 mm length of constant temperature

surrounded by a layer where the temperature is higher due to penetrating surrounding

air through the nitrogen shroud. The axial centerline temperature profile is in excellent

agreement with the flame calculations, indicating the robustness of the GRI-Mech 3.0

[16] mechanism in predicting the burning velocities of CH4/air flames and marginal

radiative heat losses in these flames. The excellent agreement between measured and

calculated axial temperature profiles was observed in all flames studied.

Page 108: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 5

103

Figure 5.1. Radial temperature profile measured in flame A at 10 mm above the burner surface.

Figure 5.2. Axial centreline temperature profile measured in flame A. Solid line and diamonds denote flame calculations and measurements, respectively.

Acetylene mole fractions, measured by Raman scattering in flame C, and shown

in figure 5.3, reach a maximum at an axial distance between 2 and 3 mm and then

decrease to ∼ 500 ppm at 9 mm, the detection limit of the current setup [15]. As can

be seen in this figure, the profile obtained with the probe is shifted approximately

1.3 mm farther downstream. A similar shift between probe and optical measurements

Page 109: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 5

104

was observed in temperature and hydroxyl profiles in other flames [17,18], and is the

result of the acceleration of the combustion products into the probe orifice [13].

Shifting the probe profile results in agreement with the Raman profiles to better than

20% (also observed in Ref. [14] at ϕ = 1.58), which substantiates the extractive probe

technique for the measurements of acetylene presented below.

Figure 5.3. Axial centerline profiles of acetylene mole fraction in methane/air flame, ϕ = 1.50 and ρv = 0.008 g/(cm2·s). Symbols denote Raman (triangles) and probe (squares with solid line) measurements; the dashed line denotes the shifted probe measurements.

At mole fractions below 500 ppm the acetylene Raman spectrum was barely

distinguishable in the noise, while the signal-to-noise ratio of the TDLAS spectrum

remained higher than 10 for mole fractions down to 100 ppm. This difference in the

limit of detectability precludes comparison of probe and Raman data in the flames

with low C2H2 mole fraction in the post-flame zone. However, due to the modest

changes in flame structure upon changing the equivalence ratio from ϕ = 1.5 to

ϕ = 1.4, we do not expect the accuracy of the probe measurements observed at

ϕ = 1.58 and ϕ = 1.5 to deteriorate substantially. The results of the extractive probe

measurements of acetylene at equivalence ratios ϕ = 1.5, 1.45 and 1.4 at different

mass fluxes are presented in figure 5.4-5.6. Consistent with the results in figure 5.3

and those presented in Ref. [14], all experimental acetylene profiles are shifted

Page 110: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 5

105

1.3 mm towards the burner surface. As can be seen from the figures, at a fixed

equivalence ratio the maximum C2H2 mole fraction depends only slightly on the mass

flux, while C2H2 oxidation in the post-flame zone increases substantially in the flames

at higher the mass flux, caused by the higher gas temperatures (given in Table 5.1). At

the same time, decreasing the equivalence ratio from ϕ = 1.5 to 1.4 decreases the

peak C2H2 mole fraction by nearly a factor of two.

Figure 5.4. Axial profiles of acetylene mole fraction in methane/air flames, ϕ = 1.5. Symbols denote probe measurements in flames A (squares), B (diamonds) and C (triangles). The dashed lines denote flame calculations with GRI-Mech 3.0, and the solid lines are the results of calculations with the increased rate coefficient for C2H2 + OH CH2CO + H discussed in the text.

Page 111: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 5

106

Figure 5.5. Axial profiles of acetylene mole fraction in methane/air flames, ϕ = 1.45. Symbols denote probe measurements in flames D (squares), E (diamonds) and F (triangles). Solid lines denote flame calculations with the increased rate coefficient for C2H2 + OH CH2CO + H discussed in the text.

Figure 5.6. Axial profiles of acetylene mole fraction in methane/air flames, ϕ = 1.40. Symbols denote probe measurements in flames G (squares), H (diamonds) and I (triangles). Solid lines denote flame calculations with the increased rate coefficient for C2H2 + OH CH2CO + H discussed in the text.

Page 112: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 5

107

In addition, we note substantial discrepancies between the measured acetylene

profiles and those obtained from the flame calculations using GRI-Mech 3.0. As can

be seen in figure 5.4, the calculations give substantially higher peak concentrations

and slower decay in the post-flame zone than those measured, well outside the 20%

differences observed between the experimental methods. The computed profiles at the

other equivalence ratios showed discrepancies similar to those presented in figure 5.4

This discrepancy has been observed previously [11], where it was attributed to the

choice of the rate coefficient of the reaction C2H2 + OH CH2CO + H used in GRI-

Mech 3.0. Following the suggestion made in Ref. [11] we increased the pre-

exponential factor of the rate coefficient to 1.7·1012 cm3/mole·s, and these results are

also presented in Figs. 5.4-5.6. The calculated acetylene profiles are now in excellent

agreement for all flames studied here. Although the limited parameter variation in the

present work precludes an unambiguous recommendation regarding increasing the

rate coefficient of this reaction, the agreement between experiment and calculations

favors this recommendation.

5.4 Conclusions

We report the measurements of acetylene in fuel-rich atmospheric-pressure

methane/air flames using spontaneous Raman and extractive probe sampling

techniques. Excepting a shift of approximately 1.3 mm, resulting from the

acceleration of the combustion products in the probe orifice, the axial Raman and

probe profiles are in very good agreement. This result validates using the extractive

probe sampling technique as a diagnostic tool for measurements of acetylene for the

conditions studied. Substantial disagreement is observed between the experimental

profiles of acetylene and those obtained from calculations based on GRI-Mech 3.0,

which predict higher acetylene concentrations and slower decay in the post-flame

zone. Increasing the pre-exponential factor in the rate coefficient for the reaction

C2H2 + OH CH2CO + H to the value of 1.7·1012 cm3/mole·s brings the calculated

acetylene profiles into excellent agreement with those derived experimentally. Further

improvement of the sensitivity of both spontaneous Raman and extractive probe

techniques will provide more information on acetylene chemistry in fuel-rich

methane-air flames. These improvements are currently in progress in our laboratory

Page 113: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 5

108

Literature

1. J. Warnatz, H. Bockhorn, A. Mozer, H. W. Wenz, Proc. Combust. Inst. 19

(1982) 197-209.

2. P. Lindstedt, Proc. Combust. Inst. 27 (1998) 269-285.

3. H. Richter, J.B. Howard, Prog. Energy Combust. Sci. 26 (4-6) (2000) 565-

608.

4. A. Dollet, Surf. Coat. Technol. 177 (2004) 245-251.

5. C. P. Fenimore, G. W. Jones, J. Chem. Phys. 41 (7) (1964) 1887-1889.

6. J. Vandooren, P. J. van Tiggelen, Proc. Combust. Inst. 16 (1977) 1133-1144.

7. D. Bittner, J. B. Howard, Proc. Combust. Inst. 19 (1982) 211-221.

8. E. W. Kaiser, J. Phys. Chem. 94 (11) (1990) 4493-4499.

9. I. T. Woods, B. S. Haynes, Combust. Sci. Technol. 87 (1-6) (1993) 199-215.

10. I. T. Woods, B. S. Haynes, Proc. Combust. Inst. 25 (1994) 909.

11. E. W. Kaiser, T. J. Wallington, M. D. Hurley, J. Platz, H. J. Curran, W. J.

Pitz, C. K. Westbrook, J. Phys. Chem. A. 104 (2000) 8194-8206.

12. J. A. Miller, C. F. Melius, Proc. Combust. Inst. 22 (1988) 1031.

13. E. L. Knuth, Combust. Flame 103 (3) (1995) 171-180.

14. A. V. Mokhov, S. Gersen, H. B. Levinsky J. Chem. Phys. Let., 403 (4-6)

(2005) 233-237.

15. A. V. Mokhov, C. E. van der Meij, H. B. Levinsky, Appl.Opt. 36 (1997)

3233-3243.

16. G. P. Smith, D. M. Golden, M. Frenklach, N.W. Moriarty, B. Eiteneer, M.

Goldenberg, C. T. Bowman, R.K. Hanson, S. Song, W.C. Gardiner, V.

Lissanski, Z. Qin, http://www.me.berkeley.edu/gri_mech/.

17. R. J. Cattolica, S. Yoon, E. L. Knuth, Combust. Sci. Technol. 28 (5-6) (1982)

225-239.

18. A. T. Hartlieb, B. Atakan, K. Kohse-Hoinghaus, Combust. Flame 121 (4) (2000) 610-624.

Page 114: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

109

Chapter 6

HCN formation and destruction in atmospheric-pressure fuel-rich premixed methane/air flames

Page 115: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

110

6.1 Introduction

To date, the vast majority of the methods developed to lower NOx emissions are

based on decreasing the flame temperature: either through flue-gas recirculation or

fuel-lean combustion. Application of these methods to obtain the lowest NOx possible

ultimately leads to deteriorating flame stability. Decreasing the temperature of fuel-

rich flames, which can be more stable than fuel-lean flames, is often not considered to

be promising as an NOx control strategy because the key initiation reaction of the

Fenimore mechanism [1]

NHCNNCH +=+ 2 (R6.1)

has a relatively low activation energy [2], and as such should be less sensitive to

flame temperature than the Zeldovich mechanism. However, experiments performed

in premixed fuel-rich natural-gas/air flames [3], showed that the measured NO mole

fractions decrease with increasing upstream heat loss, suggesting that varying the

flame temperature by altering the degree of burner stabilization can influence the NO

production in rich flames. Recent experiments at low pressure [4] have supported

these observations, and analyzed the effects of burner stabilization on Fenimore NO

formation. On the other hand, it is also possible that part of the observed decrease in

NO with flame temperature can be caused by the retardation of the conversion of the

HCN formed in reaction (R6.1), or other fixed-nitrogen species (so-called Total Fixed

Nitrogen or TFN) to NO, as suggested in [5, 6]. In this case, the “residual” TFN either

will be converted to NO in the second stage of combustion or emitted into the

environment with flue gases. To resolve this uncertainty, and to determine the

ultimate low NOx potential of fuel-rich combustion, it is essential to measure HCN in

these flames. The experimental investigations of HCN formation and destruction

performed thus far have seeded flames with bound-nitrogen additives, where the

reaction between CH and N2 is of minor importance. HCN measurements in

methane/air flames are scarce and often contradictory. For example, one study [6]

reported “as ϕ increases the maximum concentration of HCN increases initially, but

falls again in very fuel-rich flames”, while another study [7] observed a strong

increase in HCN with equivalence ratio. Moreover, Ref. [7] states that the HCN

concentration is not very dependent upon the temperature, while the reported HCN

Page 116: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

111

concentrations at T = 2560 K are almost one order of magnitude higher than those at

the same equivalence ratio (ϕ = 1.20) in Ref. [6] in the methane/air flame. Further, we

are not aware of HCN measurements in which the flame temperature was varied at

fixed equivalence ratio.

Although there are numerous experimental observations in fuel-rich flames (for

example, Ref. [8] and references therein) showing correlations between the measured

NO and CH mole fractions, as well as direct high-temperature measurements of the

rate constant of reaction (R6.1) [9,10] that reasonably agree with the results of flame

modeling, controversy about this reaction is not resolved [11]. Reaction (R6.1) as

written is “spin-forbidden”, and to reconcile theory and experiment a near unit

probability of crossing between doublet and quartet potential surfaces had to be

assumed [12]. Recently, ab initio RRKM calculations have been performed for the

reaction CH + N2 products [13,14], which showed that the reaction between CH

and N2 takes place mainly through the “spin-allowed” channel:

HNCNNCH +=+ 2 , (R6.2)

while reaction (R6.1) is not important. These calculations have been supported by

NCN detection in a low-pressure CH4/air flame [15], although the measurements were

not quantitative. It should be pointed out that theoretical results giving (R6.2) does not

contradict the experimental observations: the rate constant of the reaction (R6.1) was

determined without measurement of the products, while rapid conversion of any NCN

produced in (R6.2) to NO will preserve the correlation between the CH mole fraction

and NO formation. The experimental determination of the concentration profiles of

HCN in flames in which we expect significant Fenimore NO formation will help

resolve the uncertainty as to both the primary products of CH + N2 and the role of

HCN as a stable intermediate in NO formation.

Towards this end we measure the axial profiles of the mole fraction of HCN in

burner-stabilized rich-premixed methane/air flames at equivalence ratios ϕ = 1.3-1.5

at different flame temperatures, similar to the method followed in Chapter 5. In these

flames the HCN mole fractions are of order of a few tens of ppm. Whereas the

sensitivity of the absorption method allows measuring such low concentrations, the

strong background from the hot bands in the water absorption spectrum has frustrated

Page 117: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

112

direct in-situ HCN measurements in flames thus far. To circumvent this problem, we

follow the approach described in Chapter 4, and use microprobe sampling and

wavelength modulation absorption spectroscopy with second harmonic detection,

supplemented with CARS temperature measurements. As was done in Chapter 5 for

acetylene, the experimental observations are compared with one-dimensional flame

calculations.

6.2 Experimental

The premixed atmospheric pressure fuel-rich methane/air flames were examined

using the setup and experimental methodology described in Chapter 4, and briefly

summarized here. The flame temperature was varied by changing the mass flow

through the McKenna Products burner and measured by broadband planar BOXCARS

for nitrogen thermometry, as described elsewhere [16]. Again, the accuracy of the

CARS measurements is estimated at better than 40 K [16]. The flows of all gases

were measured by mass flow meters, while the equivalence ratio was determined by

measuring the methane volume fraction in the unburned gas/air mixture. For

calibration purposes, nitrogen doped with a known amount of HCN was flowed

through the burner instead of the methane-air mixture. In the current set-up, as

described in chapter 4, the accuracy of the measured HCN concentrations in the

calibration gases was better than 30% at mole fractions above 3ppm. Measurements

were obtained at different axial positions in the flame by moving the burner with a

precision positioner relative to the CARS laser beams and sampling probe in steps of

1 mm. The sampled gas mixtures were analyzed using WMAS with detection at the

second harmonic. Estimates of the sample cooling process showed that in the present

setup the conversion of HCN during sampling is less than 10% for all measured

flames. These estimations are supported by measurements at different suction

backpressures, which showed no significant changing in the measured HCN

concentration when varying pressure from 0.05 to 0.15 Bar.

Page 118: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

113

6.3 Results and discussion

The measurements were performed in a set of fuel-rich flames with different

equivalence ratios and mass fluxes. The flame parameters (equivalence ratios, mass

fluxes and temperatures measured at 5 mm above the burner surface) are presented in

Table 6.1.

Table 6.1

Flame ϕ ρv, g/cm2·s T, K

A 1.3 0.007 1775

B 1.3 0.011 1855

C 1.3 0.015 1942

D 1.4 0.008 1842

E 1.4 0.01 1910

F 1.4 0.014 1950

G 1.5 0.005 1763

H 1.5 0.008 1852

The temperature measurements showed that all the flames studied had a domain

with constant temperature extending at least 20 mm radially from the centerline, from

3 mm to 15 mm in the axial direction. Moreover, the measured temperatures in this

domain were in excellent agreement with the flame calculations using GRI-Mech 3.0

[17], indicating marginal radiative heat losses in these flames.

The axial HCN profiles measured in flames with equivalence ratios ϕ = 1.3, 1.4

and 1.5 are presented in figures 6.1, 6.2 and 6.3, respectively. The experimental

profiles of HCN mole fraction were shifted downstream, as in done chapter 5, where

the shift of ~1.3 mm was observed between acetylene concentrations profiles

measured by the Raman and probe techniques. Since the equivalence ratios and mass

fluxes are close to those in chapter 5, we apply the same shift (1.3 mm) to the

measured HCN profiles.

Page 119: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

114

Figure 6.1 Axial profiles of HCN mole fractions in methane air flames, ϕ = 1.3. Symbols denote probe measurements in flames A (squares), B (diamonds) and C (triangles). Lines denote flame calculations in flames A (solid), B (dashed) and C (dotted).

Figure 6.2. Axial profiles of HCN mole fractions in methane air flames, ϕ = 1.4. Symbols denote probe measurements in flames D (squares), E (diamonds) and F (triangles). Lines denote flame calculations in flames D (solid), E (dashed) and F (dotted).

Page 120: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

115

Figure 6.3. Axial profiles of HCN mole fractions in methane air flames, ϕ = 1.5. Symbols denote probe measurements in flames G (squares) and H (diamonds). Lines denote flame calculations in flames G (solid) and H (dashed).

As can be seen from figure 6.1-6.3, the measured HCN concentrations peak in

the region of the flame front, and decrease (albeit gradually) in the post-flame zone.

The maximum measured HCN mole fractions are only slightly dependent upon

equivalence ratio and mass flux, and do not exceed 15 ppm in all the flames studied

here. In flames with ϕ = 1.3 and 1.4, increasing the mass flux at fixed equivalence

ratio results in shifting the maxima towards the burner surface and accelerating the

HCN burnout. Both effects are expected based upon the reduced degree of

stabilization and the concomitant increase in flame temperature. At ϕ = 1.5 the

maximum is shifted so far downstream that the HCN burnout region is beyond the

measurement domain. The observation of 5-10 ppm of “residual” HCN downstream

of the flame front that is only slowly oxidized supports the conclusion drawn in earlier

studies of NO formation in the burnout region of rich methane flames [5]. It is

interesting to point out the strong temperature dependence of the HCN burnout. For

example, the change in residence time between flames D and F by nearly a factor of

two should lead to an increase in the HCN mole fraction at the same axial position.

However, we observe a decrease in HCN mole fraction by nearly a factor of two, most

likely caused by the relatively modest (~100 K) increase in the flame temperature.

Also interesting is the strong shift downstream of the maximum in the HCN profiles

Page 121: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

116

and the reduced oxidation rate in the post flame zone with increasing equivalence

ratio at fixed flame temperature, illustrated by comparing the flames B, D and H,

which have almost the same temperature (~1850 K) but increasing the equivalence

ratios from ϕ = 1.3 to 1.5.

Large discrepancies are observed between the measured results and those

calculated using the GRI-Mech 3.0 chemical mechanism. As can be seen from figures

6.1– 6.3, the calculations significantly overpredict the measured HCN mole fractions

for all mass fluxes at equivalence ratios ϕ = 1.3 and 1.4. At ϕ = 1.5, the calculated

HCN profiles are in surprisingly good agreement with the measurements for

ρv = 0.008 g/(cm2s), while the experimental results are underpredicted for ρv = 0.005

g/(cm2s). At the same time, GRI-Mech 3.0 predicts the qualitative trends in the

burnout region reasonably well. This suggests that adjustment of the rates of the

reactions that form and consume HCN in the flame front in GRI-Mech 3.0 would

improve the predictions substantially.

According to the GRI-Mech 3.0 scenario, HCN is formed mainly in reaction

(R6.1) between CH and N2. A substantial part of the nitrogen atoms produced in this

reaction will also be converted to HCN. First they react with CH3 to form H2CN

HCNHNCH +=+ 23 (R6.3) and then H2CN dissociates into HCN and H; here too, the HCN produced will be

oxidized to NO. Oxidation of HCN in the flame front occurs mainly by

HNCOOHCN +=+ . (R6.4) To lower the predicted peak HCN mole fraction, either the rate of formation must be

decreased or the rate of consumption increased.

To improve the agreement for the results shown above, the rate of the reaction

(R6.1) must be decreased by a factor 2.5. This reaction, being the first step of the

Fenimore mechanism, ultimately determines amount of NO that will be produced in

fuel-rich hydrocarbon flames [1,2]; changing its rate will result in large disagreement

with numerous experimental and modeling studies of the NO formation, for which

there is currently good agreement [4] en references therein). Alternatively, decreasing

the calculated HCN mole fraction in the flame front by varying the rate constant of the

Page 122: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

117

reaction (R6.4) requires increasing its rate by more than 10 times. The rate constant of

this reaction was determined both experimentally and by transition state calculations

[18-20]. Very good agreement (~30%) was observed between calculations and

measurements in the temperature region from 450 K up to 2400 K. Moreover, the rate

constant of this reaction was used as an optimization variable for the GRI-Mech 3.0

mechanism, where it was already increased by a factor 1.45. Further increase will

bring the rate constant of reaction (R6.4) far beyond the specified margins of

uncertainty.

A possible explanation for the discrepancy between measured and calculated

HCN concentrations can be found in considering the products of the reaction between

CH and N2 asserted in ref. [14]. The NCN formed in this reaction can be converted

directly to NO in the following reactions:

NOCNONCN +=+ (R6.5)

and

NONCOONCN +=+ 2 , (R6.6)

while HCN is formed in reactions

NHCNHNCN +=+ (R6.7)

and

NOHCNOHNCN +=+ . (R6.8) Reactions (R6.5)-(R6.8) are supposed to be fast [15,21], with the result that the rate of

the reaction between CH and N2 controls NO formation. At the same time, reactions

(R6.5) and (R6.6) produce NO directly while bypassing HCN formation.

Unfortunately, information on NCN kinetics is very scarce. The reactions (R6.5)-

(R6.8) were included in the modeling study of the NO reburning in a flow reactor

[21], where their rates were estimated. It should be pointed out that the reaction (R6.6)

between NCN and O2 can be important in the fuel-rich highly stabilized flames, where

oxygen is removed relatively slowly, yielding the broad flame front in which the O2

concentrations are one to two orders larger than those of O, H and OH .

Page 123: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

118

To examine the possible improvement achievable using this “NCN-route”, we

perform the calculations for flame F, for which a large discrepancy between the

measurements and numerical prediction is observed (figure 6.4). In these calculations,

we used the same rate constant for the reaction (R6.2) as for the reaction (R6.1).

Whereas this value is approximately 5 times larger than that calculated in Ref. [14],

where this reaction was first proposed, it falls within the expected accuracy of the

calculations. Recent modeling work [22] also suggests using the same rate constant

for the reaction (R6.2) as for (R6.1), and low-pressure experiments and modeling [4]

also support using this value. As can be seen from figure 6.4, using the rate constants

from [21] for the reactions consuming NCN results in decreasing the maximum

calculated HCN mole fraction to 29 ppm, 10 ppm less than that predicted by GRI-

Mech 3.0. Putting the magnitudes of the rates of all NCN removal reactions equal to

1.0·1013cm3/(mol·s) leads to a further 10 ppm reduction in the maximum HCN mole

fraction, and brings the calculated HCN profile even closer to the measured result.

Although not presented to avoid clutter in the figure, the calculations show that

varying the rates of the reactions (R6.5)-(R6.8) has only a minor influence on the NO

formation in these flames, this being determined by the rate of the initial nitrogen

fixation reaction (R6.2).

Page 124: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

119

Figure 6.4. Axial profiles of HCN mole fractions in methane air flame F. Triangles denote measured HCN concentrations. Lines denote flame calculations using GRI-Mech 3.0 mechanism with NCN removal reactions [21] (solid – unchanged rate constants, dashed – the rate constants are 1.0·1013cm3/(mol·s) for all NCN removing reactions) and dotted - GRI-Mech 3.0 without NCN removal reactions).

It should be pointed out that variation of the rate constants should be performed

very cautiously, remaining inside any expected uncertainty limits. A recent theoretical

study [23] and experiments at room temperature [24] both yielded rates of the reaction

(R6.6) between NCN and O2 that are orders of magnitude lower than that from Ref.

[21]. This large disparity dissuades us from trying optimizing the rate constants of the

reactions (R6.5)-(R6.8) based on the results presented here. In addition, it was

possible to optimize the rates to fit one experiment adequately; however, good

prediction for one experiment resulted in a poorer prediction for other experiments.

Further experimental and theoretical studies of NCN kinetics are needed for a better

understanding of NO and HCN formation in hydrocarbon flames.

Page 125: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

120

6.4 Conclusions

Axial profiles of hydrogen cyanide have been measured in laminar, atmospheric-

pressure, rich-premixed, methane/air flames at equivalence ratios ϕ = 1.3, 1.4 and 1.5.

The measurements were performed by microprobe gas sampling followed by analyses

using wavelength modulation absorption spectroscopy with second harmonic

detection. In the richest flame under investigation (ϕ = 1.5), very slow removal HCN

is observed in the post flame zone, demonstrating “residual” HCN in the post-flame

gases of fuel-rich methane/air flames. In this flame HCN concentrations of ~ 10 ppm

are measured at 7 mm above the burner surface. In practical combustion systems, this

HCN will most likely be oxidized to NO in a secondary combustion step. Decreasing

the equivalence ratio leads to faster HCN removal in the post flame zone. When

varying the flame temperature at fixed equivalence ratio no significant changes in the

HCN peak concentration is observed, while the HCN removal becomes faster with

increasing gas temperature.

Substantial disagreement is observed between the experimental profiles of HCN

and those obtained from calculations using GRI-Mech 3.0. Changing the rates of key

formation and consumption reactions showed that bringing the calculations using

GRI-Mech 3.0 into agreement with the present results can be done only at the cost of

unreasonable changes in the rates of these reactions. On the other hand, considering

NCN as a primary product of the reaction between CH and N2, based on recent

theoretical studies, allows improvement in the agreement between measured and

calculated HCN mole fractions. The lack of information on the rate constants of the

NCN reactions at high temperatures precludes unambiguous conclusions regarding

this mechanism. To provide this information we are planning to perform

measurements of CH, NO HCN and NCN in low pressure flames.

Page 126: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

121

Literature 1. C. P. Fenimore, Proc. Combust. Inst. 13 (1971) 373.

2. J. A. Miller, C. T. Bowman, Prog. Energy Combust. Sci. 15 (4) (1989) 287-

338.

3. A. V. Mokhov, H. B. Levinsky, Proc. Combust. Inst. 29 (1996) 2147-2154.

4. V. M. van Essen, A. V. Sepman, A. V. Mokkov, H. B. Levinsky., Proc.

Combust. Inst. 23 (2007) 329-337.

5. A. V. Mokhov and H. B. Levinsky., Combust. Flame 118 (1999) 733-740.

6. B. S. Haynes, D. Iverach, N. Y. Kirov, Proc. Combust. Inst. 15 (1975) 1103-

1112.

7. C. Morley, Combust. Flame 27 (2) (1976) 189-204.

8. K. Kohse-Hoinghaus, R. S. Barlow, M. Alden, E. Wolfrum, Proc. Combust.

Inst. 30 (2005) 89-123.

9. D. Lindackers, M. Burmeister, P. Roth, Proc. Combust. Inst. 23 (1990) 251-

257.

10. A. J. Dean, R. K. Hanson, C.T. Bowman, Proc. Combust. Inst. 23 (1990)

259-265.

11. J. A. Miller, M.J. Pilling, E. Troe, Proc. Combust. Inst. 30 (2005) 43-88.

12. J. A. Miller, S. P. Walch, Int. J. Chem. Kinet. 29 (4) (1997) 253-259.

13. L. V. Moskaleva, W. S. Xia, M. C. Lin, Chem. Phys. Lett. 331 (2-4) (2000)

269-277.

14. L. V. Moskaleva, M. C. Lin, Proc. Combust. Inst. 28 (2000) 2393-2401.

15. G. P. Smith, Chem. Phys. Lett. 367 (5-6) (2003) 541-548.

16. A. V. Mokhov, C. E. van der Meij, H. B. Levinsky, Appl.Opt. 36 (1997)

3233-3243.

17. G.P. Smith, D.M. Golden, M. Frenklach, N.W. Moriarty, B. Eiteneer, M.

Goldenberg, C.T. Bowman, R. Hanson, S. Song, W.C. Gardiner, V.

Lissanski, Z. Qin, http://www.me.berkeley.edu/gri_mech/.

18. P. Roth, R. Lohr, H. D. Hermanns, Berichte der Bunsen-Gesellschaft-

Physical Chemistry Chemical Physics 84 (9) (1980) 835-840.

19. M. Louge, R. K. Hanson, Proc. Combust. Inst. 20 (1985) 665-675.

20. R. A. Perry, C. F. Melius, Proc. Combust. Inst. 20 (1984) 639-646.

Page 127: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 6

122

21. P. Glarborg, M.U. Alzueta, K. Dam-Johansen, J. A. Miller, Combust. Flame

115 (1-2) (1998) 1-27.

22. A. El bakali, L. Pillier, P. Desgroux, B. Lefort, L. Gasnot, J. F. Pauwels, I. da

Costa, Fuel 85 (2006) 896-909.

23. R. S. Zhu, M. C. Lin, Int. J. Chem. Kinet. 37 (10) (2005) 593-598.

24. R. E. Baren, J. F. Hershberger, J. Phys. Chem. A 106 (46) (2002) 11093-

11097.

Page 128: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

123

Chapter 7

The effect of hydrogen addition to rich stabilized methane/air flames

Page 129: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

124

7.1 Introduction

Stringent emission regulations for greenhouse gases and the drive to conserve

the finite supplies of fossil fuels have resulted in increased interest in the possible

addition of sustainable hydrogen to natural gas. In spite of any potential advantage to

hydrogen addition in this regard, a negative effect on other aspects of combustion

performance, such as increased pollutant emissions (NOx, soot), must be weighed in

the overall considerations. Recently, several studies on the effect of hydrogen addition

to hydrocarbons have been conducted regarding extinction limits [1], burning

velocities [2-4] and engine performance [5]. However, very few flame structure

studies have been performed on hydrogen-hydrocarbon mixtures, and all of these

pertained to non-premixed flames. For example, the effect of hydrogen was studied in

a hydrogen-natural-gas composite fuel in turbulent jet flames [6], and showed that

increasing the hydrogen concentration resulted in an increase in soot, CO and NOx

formation. Another study [7] added natural gas and propane to coflow hydrogen

diffusion flames; a decrease in soot formation and an increase in NOx with increasing

hydrogen content were found. In the same type of flames, intermediate radicals were

measured [8,9] and a decrease in CH radical concentration was observed with

increasing hydrogen content. In these studies, the increase in NOx emission was

ascribed to the increasing flame temperature with hydrogen content, which promotes

thermal NOx production. However, the measured decrease in CH concentration with

hydrogen content suggests that hydrogen addition to hydrocarbon flames reduces the

NOx emissions contributed from the Fenimore mechanism. The analysis of the

chemistry in the coflow diffusion flame studies described above suffers from the

complication of multidimensional transport and the fact that the difference in diffusive

behavior between the relative heavy hydrocarbon fuels and hydrogen is very large.

For this reason, measurements of key intermediates species in pollutant formation

using premixed one-dimensional flames can yield a more unambiguous contribution

to understanding the consequences of hydrogen addition for the chemistry of pollutant

formation. To date, no quantitative studies have been reported on the effect of

hydrogen addition on pollutant formation in one-dimensional flames.

One of the key intermediate species in the formation of soot in rich hydrocarbon

flames is acetylene [10] and understanding the fate of this species is essential to

understand soot inception. Also, acetylene is a precursor of the CH radical [11], which

Page 130: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

125

is an important intermediate in Fenimore NO formation in hydrocarbon flames [12].

As discussed in chapter 6, HCN is an important intermediate in the Fenimore

mechanism as well. Moreover, due to the lack of oxygen radicals in fuel-rich flames

the conversion of HCN to NO is slow and the poisonous HCN can survive in the post-

flame gases (chapter 6). As mentioned in chapter 5 and 6, and reference therein, both

acetylene and HCN have been investigated in flat, atmospheric-pressure premixed

hydrocarbon flames. However, to our knowledge no quantitative studies have been

reported of the effects of hydrogen addition on the formation and destruction of C2H2

and HCN in premixed CH4/air flames. Since there are substantial concentrations of

HCN and C2H2 in rich premixed flames, we study these effects in flames similar to

those discussed in chapter 5 and 6. Towards this end we measured the profiles of

C2H2 and HCN in rich-premixed H2/CH4/air flames at atmospheric pressure. The

technique used is microprobe gas sampling followed by analyses using tunable diode

laser absorption spectroscopy and wavelength modulation absorption spectroscopy

with second harmonic detection (chapter 4). In addition, the experimental

observations are compared with one-dimensional flame calculations.

Page 131: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

126

7.2 Experimental

The flat, atmospheric-pressure premixed flames of CH4/H2/air have been

stabilized on a 6-cm diameter flat-flame water-cooled Mc-Kenna Products burner and

surrounded by a coflow of nitrogen to prevent from mixing with ambient air (chapter

4). The experimental conditions are summarized in table 7.1.

Table. 7.1. Flame parameters

Flame ϕ H2 (%) ρv, g/cm2·s) T, K (GRI-3.0)

A 1.5 0 0.005 1760

B 1.5 23 0.005 1740

C 1.3 0 0.005 1763

D 1.3 23 0.005 1743

E 1.4 0 0.008 1854

F 1.4 20 0.008 1833

G 1.4 50 0.010 1834

H 1.4 50 0.008 1797

I 1.5 0 0.008 1848

J 1.5 20 0.008 1829

K 1.5 50 0.008 1796

Measurements were obtained at different vertical positions in the flame by

moving the burner with a precision positioner up to a distance of 10 mm in steps of

1 mm. As was done in the previous chapters, all measured HCN and C2H2

concentration profiles are shifted 1.3 mm upstream to correct for the probe distortion.

The methods for obtaining the C2H2 and HCN mole fractions, via tunable diode laser

absorption spectroscopy are described in detail in chapter 4.

Given the excellent predictive power of Chemkin II [13] with GRI-Mech 3.0

[14] for predicting the flame temperature observed in chapter 5 for methane/air

flames, and in other studies in methane/air [15] and hydrogen/air flames [16],

calculated flame temperatures were used in this study (Table 7.1). Calculations were

Page 132: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

127

performed by using the measured mass flux through the burner surface as an input

parameter.

7.3 Results and discussion

The addition of hydrogen to the unburned methane/air mixture changes the

flame properties. The laminar flame velocity of pure hydrogen is 8 times higher that

that of pure methane [4]; we thus expect that the flame velocity of methane to increase

substantially upon hydrogen addition. When the flame is stabilized above the burner

surface, we expect hydrogen addition to cause a significant temperature reduction; at

constant mass flux through the burner, the higher flame velocity should result in more

heat transfer to the burner. To illustrate this, figure 7.1 presents the calculated flame

temperature as function of the mass flux for methane and a methane/hydrogen

mixture. As can be seen, replacement of 50% methane by hydrogen results, at

constant mass flux, in a substantial decrease in the calculated flame temperature,

caused by increased heat transfer to the burner.

Figure 7.1. Calculated flame temperature as function of the mass flux for a pure methane (solid line) and a methane/hydrogen, 50/50 mixture (dashed line), both having an equivalence ratio of ϕ=1.5.

Page 133: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

128

7.3.1 HCN profiles To study the effect of hydrogen addition to fuel-rich CH4/air flames on the

formation and consumption of HCN, several flames with different equivalence ratios

(ϕ=1,3, 1.4 and 1.5) and hydrogen concentrations in the fuel mixture (0%, 20% and

50% by volume) have been studied (see Table 7.1). For the flames at ϕ=1.5 (flames I-

K) and ϕ=1.4 (flames E, F, H) the mass flux is kept constant with different hydrogen

content in the mixture. As expected from figure 7.1, hydrogen addition to the

methane/air mixtures while keeping the mass flux constant decreases the flame

temperature (I-K and E, H) slightly (see also Table 7.1). For the flames with

equivalence ratio ϕ=1.4 (flames E-G), the temperature was kept more or less constant

(±20K) by increasing the mass flux when more hydrogen was added.

The experimental profiles of HCN mole fraction in Figs. 7.2 and 7.3, for ϕ=1.4

and ϕ=1.5, respectively, show that as the fraction of hydrogen in the fuel mixture

increases, the HCN mole fraction decreases substantially, well outside the

measurement uncertainty. For example, increasing the hydrogen content in the fuel

from 0 to 50% lowers the HCN mole fraction by more than a factor of two, at both

equivalence ratios. According to reaction (R6.1) in Chapter 6, HCN should be linearly

related to the CH and N2 concentration. Furthermore, one could make the simple

“naïve” assumption that CH is directly proportional to the CH4 concentration in the

fuel/air mixture. Following this simple line of thought, replacing 20% (v/v) methane

by hydrogen in the fuel results in a reduction, by “dilution”, in the CH4 mole fraction

by ∼7% and ∼2% in the N2 mole fraction in the fuel-air mixture; taken together this

would in turn decreases the HCN concentration via reaction R6.1 by ∼10%. When

increasing the hydrogen fraction to 50% a HCN decrease of ∼30% could thus be

expected. That a factor of 2 decrease in HCN is observed for a hydrogen fraction of

50%, suggests an additional effect, besides dilution, on flame structure. Figure 7.2

shows, for the flames with 50% hydrogen addition (flames D and J), that lowering the

temperature by ∼ 60 K does not affect the HCN peak concentration significantly, as

also seen for the “pure” methane/air flames in chapter 6. Interesting to note in figures

7.2 and 7.3 is that at constant mass flux the HCN peak concentrations appears to shift

towards the burner surface as the fraction of hydrogen in the fuel increases from 0 to

50%, indicative of the higher degree of burner stabilisation caused by the increased

burning velocity.

Page 134: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

129

Figure 7.2. Axial profiles of HCN mole fraction in CH4/air flames and CH4/H2/air flames, ϕ = 1.4. Symbols denote probe measurements in flames E (squares), F (diamonds), G (circles) and H (triangles). The lines denote flame calculations using GRI-Mech 3.0 in flames E (dashed), F (dotted) G (bold solid) and H (thin solid).

Figure 7.3. Axial profiles of HCN mole fraction in CH4/air flames and CH4/H2/air flames, ϕ = 1.5. Symbols denote probe measurements in flames I (diamonds), J (squares) and K (triangles). The solid lines denote flame calculations using GRI-Mech 3.0 in flames I (dashed), J (solid) and K (dotted).

Page 135: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

130

Comparison of the calculated and measured HCN concentration profiles at

ϕ=1.4 (figure 7.1) shows for all flames substantial overprediction of the HCN peak

concentration and significant slower HCN decay relative to the experimental results,

similar to those observed in Chapter 6 for the pure methane fuel. Moreover, whereas a

reduction in the measured HCN peak concentration at ϕ=1.4 is related to the amount

of hydrogen addition, with only a modest effect of temperature, the calculations

suggest that the HCN reduction is mainly related to the flame temperature and not due

to a “hydrogen” effect. For example, increasing the hydrogen content from 20%

(flame F) to 50% (flame G) at constant flame temperature does not reduce the

maximum calculated HCN concentration, but lowering the flame temperature while

keeping 50% hydrogen content in the fuel (flame H) results in a substantial reduction

in the calculated HCN peak concentration. Although it is tempting to further interpret

the calculations, we note that these observations are in contradiction with the

measurements. It is thus not prudent to pursue the analysis using GRI-Mech 3.0. A

possible explanation for the large quantitative and qualitative discrepancies between

calculated and measured HCN peak concentrations observed at ϕ=1.4 (figure 7.2) is

that in GRI Mech 3.0 the temperature sensitive reaction CH+N2=HCN+N (R6.1,

chapter 6) is the main source of formation of HCN, while as discussed in Chapter 6

recent theoretical studies [17,18] point to NCN and not HCN as the primary product

of this reaction.

In contrast to the poor agreement at ϕ=1.4, figure 7.3 shows for the flames

studied at ϕ=1.5 moderately good agreement between measured and calculated HCN

profiles. However, as observed above, the calculations are very sensitive to burner

stabilization, as reflected in the flame temperature, and we ascribe the “good”

agreement with the trend observed here for ϕ=1.5 as a spurious effect of the changes

in flame temperature. This “temperature” effect is even more dramatically illustrated

in figure 6.3; where the experimental profiles differ only modestly for a change in

flame temperature of ~100 K, while the calculations change by a factor of 3. Due to

the significant quantitative discrepancies between calculated and measured HCN

profiles observed at ϕ=1.4 (figure 7.2), and for the spurious effect of temperature

observed at ϕ=1.5 in Chapter 6 (figure 6.3), and suspected in figure 7.3, we refrain

from the further mechanistic analyse the effect of hydrogen on the HCN mole

fraction. Based on the results reported here, a reanalysis of the chemical mechanism is

Page 136: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

131

deemed necessary. To do so, detailed information on the NCN kinetics is must be

obtained. We hope that the experimental data presented here will be useful in

performing this task.

7.3.2 C2H2 profiles The effect of replacing 23% methane by hydrogen, at constant mass flux, on the

C2H2 profiles is illustrated in figure 7.4. The results show for equivalence ratios 1.3

and 1.5 (Flames A-D in Table 1) that H2 addition only slightly decreases the measured

peak C2H2 mole fraction. This small reduction of C2H2 is essentially the same as the

∼15% decrease “naively” expected from dilution when 23% methane is replaced by

hydrogen, here too making the simple assumption that C2H2 formation is directly

proportional to the concentration of hydrocarbons in the fuel. No substantial

difference in the measured C2H2 concentration is observed in the post flame zone for

the flames with (B, D) and without hydrogen addition (A, C).

In addition, the predicted C2H2 profiles obtained using GRI-Mech 3.0 are compared

with the measurements. As can be seen from figure 7.4, the calculated C2H2 profile

(dashed line) shows substantially higher peak concentrations and slower decay in the

post flame zone for flame D. Although not presented, the computed C2H2 profiles for

the other flames shown in figure 7.4 show similar discrepancies, as expected from the

results presented in Chapter 5 and in Ref. [19]. Replacing the rate coefficient of the

reaction C2H2 + OH → CH2CO +H used in the GRI 3.0 mechanism by the expression

4.87 x 1013exp(-12000cal/RT) cm3mol-1s-1 recommended in chapter 5 yields very

good agreement between measurements and calculations for all flames studied, as can

be seen in the figure.

Page 137: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

132

Figure 7.4. Axial profiles of HCN mole fraction in CH4/air flames and CH4/H2/air flames, ϕ = 1.5 and 1.3. Symbols denote probe measurements in flames A (circles), B (squares), C (diamonds) and D (triangles). The dashed line denotes flame the calculation with GRI-3.0, and other lines, flame A (bold solid lines), flame B (thin solid line), flame C (bold doted line) and flame D (thin dotted line) are the result of calculations with the increased rate coefficient for C2H2+OH→CH2CO+H.

7.4 Conclusion

In this chapter we examined the effect of hydrogen addition on the formation

and destruction of HCN and C2H2 in rich premixed CH4/air mixtures. The HCN

measurements at equivalence ratios ϕ=1.4 and 1.5 show that the HCN mole fraction

decreases substantially with increasing hydrogen content in the fuel mixture. This

decrease significantly exceeds the reduction in HCN expected from the dilution of the

hydrocarbon fuel and nitrogen when hydrogen is added to the mixture, suggesting that

addition of hydrogen affects the flame structure related to the formation of HCN. In

contradiction to the measurements, calculations, using GRI-Mech 3.0 show at ϕ=1.4

and constant flame temperature no substantial reduction in the HCN peak

concentration when hydrogen is added to the mixture, while a reduction in flame

temperature results in a substantial decrease in the calculated HCN peak concentration

at the same hydrogen fraction in the fuel (50% v/v). Moreover, calculations predict

substantially higher HCN peak mole fractions and slower oxidation in the post flame

Page 138: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

133

zone as compared to the measurements for all flames at ϕ=1.4. Increasing the

hydrogen content in the fuel for the flames at ϕ=1.5 results in roughly the same

systematic decrease in the calculated HCN concentration as measured. However, this

observed reduction is probably caused by the reduction in temperature caused by

increased stabilization, which is not supported by the measurements presented in this

thesis, and not due to a “hydrogen effect”.

The C2H2 measurements show that the addition of 23% hydrogen results in only

a marginal reduction of the C2H2 concentration for the flames measured at ϕ=1.3 and

1.5, suggesting that hydrogen addition does not have any significant effect on the

flame processes responsible for C2H2 formation/consumption in CH4/air flames.

Comparison between calculated and measured profiles shows significant

overprediction of the maximum C2H2 concentration and a much slower predicted

decay of C2H2 in the post flame zone. As expected from Chapter 5, replacing the rate

coefficient of the reaction C2H2 + OH → CH2CO +H used in the GRI-Mech 3.0

mechanism by 4.87 x 1013exp(-12000cal/RT) cm3mol-1s-1 resulted in good agreement

between measured and calculated C2H2 profiles for all flames studied.

Page 139: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

134

7.5 Literature

1. I. Wierzba , W. Wang , Int. J. hydrogen energy 31 (2006) 485-489.

2. M. Ibas , A. P. Crayford, I. Yilmaz , P. J. Bowen, N. Syred, int. J. hydrogen

energy 31 (2006) 1768-1779.

3. F. Halter, C. Chauveau, N. Djebaili-Chaumeix, I. Gokalp , Proc.Combust.

Inst. 30 (2005) 201-208.

4. B. E. Milton, J.C. Keck, Combust. Flame 58 (1984) 13-22.

5. S. O. Bade Shrestha, G. A. Karim, Int. J. Hydrogen Energy 24 (1999) 577-

586.

6. F. Cozzi A. Coghe , Int. J. hydrogen Energy 31 (2006) 669-677.

7. A. R. Choudhui, S. R. Gollahalli, Int. J. Hydrogen Energy 25 (2000) 451-

462.

8. A. R. Choudhui, S. R. Gollahalli, Int. J. Hydrogen Energy 25 (2000) 1119-

1127.

9. A. R. Choudhui , S. R. Gollahalli, Int .J. Hydrogen Energy 29 (2004)1293-

1302.

10. J. Warnatz, H. Bockhorn, A. Mozer, H.W. Wenz, Proc. Combust. Inst 19

(1982) 197-209.

11. J. Warnatz, Ber. Bunsenges. Phys.Chem 87 (1983) 1008-1022

12. C. P. Fenimore, Proc. Combust. Inst. 13 (1971) 373-379.

13. R. J. Kee, F. M. Rupley, J. A. Miller, CHEMKIN II: A Fortran Chemical

Kinetics Package for the Analysis of Gas-Phase Chemical Kinetics., Sandia

National Laboratories, (1989).

14. G. P. Smith, D. M. Golden, M. Frenklach, N.W. Moriarty, B. Eiteneer, M.

Goldenberg, C. T. Bowman, R. K. Hanson, S. Song, W. C. Gardiner, V.

Lissanski, Z. Qin, http://www.me.berkeley.edu/gri_mech/.

15. A. V. Sepman, A. V. Mokhov, H. B. Levinsky, Proc. Combust. Inst. 29

(2002) 2187-2194.

16. A. V. Sepman, Effect of burner stabilization on nitric oxide formation and

destruction in atmospheric pressure flat premixed flames, Ph.D. Thesis,

RUG, 2006 (ISBN 90-367-2702-2).

17. L.V. Moskaleva, W.S. Xia, M.C. Lin, Chem. Phys. Let. 331 (2-4) (2000)

269-277.

Page 140: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Chapter 7

135

18. L.V. Moskaleva, M.C. Lin, Proc. Combust. Inst. 28 (2000) 2393-2401.

19. E. W. Kaiser, T. J. Wallington, M. D. Hurley, J. Platz, H.J. Curran, W. J.

Pitz, C. K. Westbrook, J. Phys. Chem. A., 104 (2000) 8194-8206.

Page 141: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Summary

136

Increasingly stringent regulations regarding CO2 emissions, and the expectation

that fossil fuel reserves will be exhausted within this century, have led to interest in

the use of admixtures of hydrogen in natural gas as an alternative fuel in combustion

devices. Combustion equipment is generally tuned for local fuel used, and clearly a

change in fuel must not lead to deterioration in performance. Since the combustion

properties of hydrogen differ in many respects from those of natural gas (Chapter 1),

there are concerns regarding the possible negative response of combustion equipment

such as gas engines, burners and turbines when fuelled with hydrogen-enriched

natural gas. For example, the presence of hydrogen might increase pollutant emissions

from combustion devices and cause knock (uncontrolled ignition) in gas engines. To

understand these practical consequences properly, it is necessary to study the changes

in the underlying physical and chemical processes that are responsible for changes in

combustion behavior when hydrogen is added. The autoignition properties presented

in this thesis provide new insight into the ignition characteristics of methane,

hydrogen and methane/hydrogen fuel mixtures under conditions relevant to knock in

gas engines. In addition, the spatial profiles of C2H2 and HCN (important precursors

of soot and NOx, respectively) measured in atmospheric-pressure, one-dimensional

CH4/air and CH4/H2/air flames provide insight into changes in pollutant formation

upon hydrogen addition. To test the accuracy of different chemical mechanisms,

which could be used to predict the combustion behaviour of natural gas/hydrogen

mixtures, the measurements presented in this thesis are compared with the results of

numerical simulations.

To study autoignition under strictly controlled experimental conditions relevant

to gas engines, a Rapid Compression Machine (RCM) was constructed in our

laboratory based upon a design from MIT (Chapter 2). Test results showed that the

RCM is able to compress the combustible gas-air mixture to final pressures up to ∼70

bar and temperatures up to ∼1100 K, where the majority of the pressure rise in the

compression period takes place in a very short time (<3 ms). The temperature of the

compressed mixture is calculated from the measured pressure by using the isentropic

relations of an ideal gas, the uncertainty in the peak temperature after compression is

better than ±3.5 K in the range of pressures of interest (10-70 bar).

Chapter 3 presents the experimental study of autoignition delay times of

methane/hydrogen mixtures at high pressure (10-70 bar) and moderate temperatures

Page 142: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Summary

137

(960-1100 K). Under stoichiometric conditions the experimental results show that

replacing methane by hydrogen results in a reduction in the measured ignition delay

time. At moderately low concentrations of hydrogen (≤ 20%) only a weak effect on

the measured ignition time is observed, while at 50% hydrogen content in the fuel a

substantial reduction in ignition delay time is seen under all measured conditions.

Moreover, the measurements show that the effects of hydrogen in promoting ignition

increases with temperature and decreases with pressure. Experimental results for 50%

hydrogen in the fuel at equivalence ratio ϕ = 0.5 are essentially identical to those at

ϕ = 1.0. These results suggest that the adverse affects of hydrogen addition to natural

gas on engine knock may be limited for hydrogen fractions of only a few tens of

percent. Comparison between measured and calculated ignition delay times shows

very good agreement for all fuel mixtures using the proposed mechanism of Petersen

et al. (E. L. Petersen, D. M. Kalitan, S. Simmons, G. Bourgue, H. J. Curran and J. M.

Simmie, Proc. Combust. Inst .31 (2007) 447-454.)

Chapter 4 describes the experimental protocols for concentration measurements

of HCN and C2H2 in one-dimensional flames using extractive probe sampling

followed by analysis using tunable diode laser absorption spectroscopy (TDLAS) at

∼1.5 μm. The calibration procedure for acetylene is performed by measuring the

absorption coefficient in a gas sample containing a known concentration of acetylene

under the same experimental conditions as those existing for the flame samples. At

mole fractions above 1000 ppm, the accuracy of the measured C2H2 is ∼5%;

decreasing C2H2 mole fraction results in deteriorating accuracy, to 15% at 100 ppm.

The same calibration procedure is performed for the HCN measurements. However, to

increase the sensitivity wavelength modulation absorption spectroscopy (WMAS)

with second harmonic detection is used. The accuracy of the measured HCN mole

fraction in the sampled flame gases is better than 30% at concentrations above 3 ppm.

Before examining the effects of hydrogen addition on the formation and consumption

of these species, their profiles are first measured and analyzed in flames using pure

methane as fuel.

Chapter 5 presents measurements of C2H2 concentration profiles in one-

dimensional atmospheric-pressure rich premixed methane/air flames using

spontaneous Raman scattering and an extractive probe sampling technique (Chapter

4). Excepting a shift of approximately 1.3 mm, resulting from the acceleration of the

Page 143: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Summary

138

combustion products in the probe orifice, the axial Raman and probe profiles are in

very good agreement. The measurements show that changing the equivalence ratio

from ϕ = 1.4 to 1.5 results in an increase of the peak C2H2 mole fraction by nearly a

factor two. At fixed equivalence ratio, the maximum C2H2 mole fraction depends only

slightly on the flame temperature, while the C2H2 oxidation in the post flame zone

increases substantially in the flames with increasing flame temperature. Comparison

of measured C2H2 profiles with those calculated, using the GRI-Mech 3.0 chemical

mechanism, shows a much faster post-flame decay in the experimental results.

Increasing the pre-exponential factor in the rate coefficient of reaction

C2H2 + OH → CH2CO + H to 1.7 x 1012 cm3mol-1s-1 in the range 1760 – 1850 K

yields excellent agreement between computed and experimental results.

In Chapter 6 the formation and consumption of HCN in fuel-rich atmospheric

pressure methane/air flames is discussed. Towards this end, axial HCN and

temperature profiles have been measured at equivalence ratios ϕ = 1.3, 1.4 and 1.5.

For the richest flame studied (ϕ = 1.5) very slow oxidation of HCN in the post flame

zone is observed, demonstrating “residual” HCN in the post flame gases of fuel-rich

methane/air flames. The HCN measurements show that increasing the flame

temperature at fixed equivalence ratio does not result in significant changes in the

HCN peak mole fraction, while the HCN oxidation in the post-flame gases increases

substantially. Decreasing the equivalence ratio leads to faster HCN oxidation in the

post flame zone. Large discrepancies are observed between measured and calculated

HCN profiles using GRI-Mech 3.0. Attempts to bring the calculations using GRI-

Mech 3.0 into agreement with the experimental observations by changing the rates of

key formation and consumption reactions within the uncertainties in the literature

were unsuccessful. Consideration of NCN as a primary product of the reaction

between CH and N2, based on recent theoretical studies, allows improvement in the

agreement between measured and calculated HCN concentrations. However, the lack

of information on the rate constants of the NCN reactions at high temperatures

precludes unambiguous conclusions regarding this mechanism.

Chapter 7 is an extension of Chapters 5 and 6 and examines the effect of

hydrogen addition on the formation and consumption of HCN and C2H2 in fuel-rich

stabilized methane/air flames. The HCN measurements at ϕ = 1.4 and 1.5 show that

increasing the hydrogen fraction in the mixture at constant flame temperature results

Page 144: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Summary

139

in a substantial decrease in the HCN mole fraction. This decrease in HCN

significantly exceeds the reduction of HCN caused by the dilution of hydrocarbon fuel

when hydrogen is added to the mixture, indicating that hydrogen addition affects the

chemistry related to the formation of HCN. As observed for pure methane flames

(Chapter 6) at ϕ = 1.4, the calculated HCN profiles using GRI-Mech 3.0 predict

significantly higher HCN peak mole fractions and substantially slower decay in

comparison to the measurements for all flames studied. Moreover, contrary to the

experimental observations, the calculations show no substantial changes in the

calculated peak HCN mole fraction when adding hydrogen to the fuel mixture and a

strong reduction in the HCN mole fraction with decreasing flame temperature at

constant hydrogen fraction. Good agreement between calculations and measurements

is found for the flames at ϕ = 1.5. However, the reduction in the calculated HCN mole

fraction when hydrogen is added to the flame is probably the result of the computed

reduction in flame temperature and not due to a “hydrogen” effect on the chemistry.

No significant changes are observed in the measured C2H2 mole fractions for the

flames with equivalence ratio ϕ=1.3 and 1.5 when hydrogen is added. This suggests

that hydrogen addition does not have a significant effect on the chemistry responsible

for C2H2 in CH4/air flames. Replacing the rate coefficient of the reaction C2H2 + OH

→ CH2CO +H used in the GRI 3.0 mechanism by the rate recommended in Chapter 6

resulted in good agreement between measured and calculated C2H2 profiles for all

flames studied.

Page 145: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Samenvatting

140

De steeds strengere eisen ten aanzien van CO2 emissies en de verwachting dat nog

binnen deze eeuw de reserves van fossiele brandstof uitgeput zullen zijn, heeft geleid

tot toenemende belangstelling in het gebruik van aardgas/waterstof mengsels als

alternatieve brandstof in verbrandingsapparatuur. Deze apparatuur is gewoonlijk

afgesteld voor de locale brandstof, en het zal duidelijk zijn dat veranderingen in

brandstofsamenstelling niet tot verslechterd gedrag mag leiden. Doordat de

verbrandingseigenschappen van waterstof sterk verschillen ten opzichte van aardgas

(Hoofdstuk 1), is de vraag in hoeverre het gedrag van verbrandingsapparatuur zoals

gasmotoren, branders en turbines (negatief) beïnvloed zal worden door het aardgas

met waterstof te verrijken. De aanwezigheid van waterstof kan bijvoorbeeld de

vorming van milieuschadelijke emissies bevorderen bij verbrandingsapparatuur en

leiden tot klopverschijnselen (ongecontroleerde ontsteking) in gasmotoren. Om deze

praktische consequenties adequaat te doorgronden is het essentieel de veranderingen

in de onderliggende fysische en chemische processen die verantwoordelijk zijn voor

het verbrandingsgedrag te onderzoeken, wanneer waterstof wordt toegevoegd. De

zelfontstekingseigenschappen vermeld in dit proefschrift geven nieuwe inzichten in

het ontstekingsgedrag van methaan, waterstof en methaan/waterstof

brandstofmengsels, onder condities relevant voor klopverschijnselen in gasmotoren.

Daarnaast geven de ruimtelijke verdelingen van de molfracties van C2H2 en HCN,

belangrijke tussencomponenten in de vorming van respectievelijk roet en NOx,

gemeten in één-dimensionale CH4/lucht en CH4/H2/lucht vlammen bij atmosferische

druk inzicht in veranderingen in de vorming van milieuschadelijke stoffen als gevolg

van waterstof toevoeging. De metingen die in dit proefschrift worden gepresenteerd

worden ook gebruikt om de nauwkeurigheid te testen van verschillende chemische

mechanismen die gebruikt zouden kunnen worden voor het voorspellen van het

verbrandingsgedrag van waterstof/aardgas mengsels.

Op basis van een ontwerp van MIT is in ons lab een Rapid Compression

Machine (RCM) gebouwd, waarmee onder gecontroleerde omstandigheden die

relevant zijn voor gasmotoren zelfontsteking kan worden onderzocht (Hoofdstuk 2).

Resultaten tonen aan dat de RCM een brandbaar mengsel tot drukken tot ~70 bar en

temperaturen tot ∼1100 K comprimeert, waarbij het merendeel van deze

drukverhoging plaatsvindt in een zeer korte tijd (<3 ms). De temperatuur van het

gecomprimeerde mengsel is berekend aan de hand van de gemeten druk, door gebruik

te maken van de isentropische relaties van een ideaal gas. Binnen het bereik van de

Page 146: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Samenvatting

141

gemeten drukken (10-70 bar) is de onzekerheid in de berekende piektemperaturen na

compressie kleiner dan ± 3.5 K.

Hoofdstuk 3 presenteert de experimentele studie van zelfontstekingstijden van

methaan/waterstof mengsels onder hoge drukken (10-70 bar) en gematigde

temperaturen (960-1100K). Onder stoichiometrische condities tonen de metingen aan

dat het vervangen van methaan door waterstof resulteert in een verlaging van de

gemeten ontstekingstijd. Toevoeging van lage concentraties waterstof (<20%) laat

slechts een bescheiden effect op de gemeten ontstekingstijd zien, terwijl bij de

mengsels met 50% waterstof in de brandstof een substantiële verlaging in de

ontstekingstijd wordt waargenomen. Tevens volgt uit de metingen dat het effect van

waterstof in het bevorderen van ontsteking toeneemt met toenemende temperatuur en

afneemt bij toenemende druk. De experimentele resultaten verkregen bij 50%

waterstof in de brandstof en een equivalentieverhouding ϕ = 0.5 zijn nagenoeg

identiek aan de resultaten verkregen bij ϕ = 1.0. De resultaten suggereren dat de

toevoeging van slechts enkele tientallen procent waterstof aan aardgas waarschijnlijk

een beperkte invloed zal hebben op klopverschijnselen in gasmotoren. De vergelijking

van de gemeten en berekende ontstekingstijden met het chemische mechanisme van

Petersen e.a (E. L. Petersen, D. M. Kalitan, S. Simmons, G. Bourgue, H. J. Curran and

J. M. Simmie, Proc. Combust. Inst. 31 (2007) 447-454.) laten voor alle

brandstofmengsels zeer goede overeenkomsten zien.

Hoofdstuk 4 beschrijft de meetprotocollen voor het bepalen van HCN- en C2H2-

concentraties in één-dimensionale vlammen. Hierbij worden monsters van de hete

gassen met behulp van een afzuigprobe genomen en geanalyseerd met behulp van

tunable diode laser absortption spectroscopy (TDLAS) bij 1.5 μm. De

calibratieprocedure voor acetyleen wordt uitgevoerd door de absorptiecoëfficiënt te

bepalen van een gasmonster met een bekende concentratie van acetyleen onder

dezelfde condities die gelden voor de vlammonsters. Voor molfracties groter dan

1000 ppm is de nauwkeurigheid van de gemeten C2H2 ca. 5%; de nauwkeurigheid

neemt af bij afnemende molfractie, oplopend tot 15% bij 100 ppm acetyleen. Voor de

HCN-metingen is dezelfde calibratie methode toegepast als voor C2H2. Echter, voor

het verhogen van de gevoeligheid is wavelength modulation absorption spectroscopy

(WMAS) toegepast, met detectie op de tweede harmonische frequentie. De

Page 147: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Samenvatting

142

nauwkeurigheid van de gemeten HCN molfractie in de monstergassen is beter dan

30% voor molfracties >3 ppm.

Hoofdstuk 5 presenteert metingen van de profielen van C2H2-molfracties in

één-dimensionale brandstofrijke voorgemengde methaan/lucht vlammen bij

atmosferische druk met behulp van zowel spontane Ramanverstrooiing als de

afzuigprobe-techniek (Hoofdstuk 4). Behalve een verschuiving van ongeveer 1.3 mm,

als gevolg van de versnelling van verbrandingsproducten in de probe orifice, tonen de

Raman- en probe- profielen zeer goede overeenkomst. Uit de meetresultaten volgt dat

verandering van de equivalentieverhouding van ϕ = 1.4 naar 1.5 resulteert in een

verhoging van de piekconcentratie van bijna een factor twee. Bij constante

equivalentieverhouding is de C2H2 piekmolfractie slechts weinig afhankelijk van de

vlamtemperatuur, terwijl de C2H2-oxidatie in de postvlam-zone substantieel toeneemt

voor de vlammen met toenemende vlamtemperatuur. Vergelijking van de gemeten en

de op basis van het GRI 3.0 chemische mechanisme berekende C2H2-profielen laat

een veel snellere afname in C2H2-molfractie in de metingen zien. Verhogen van de

pre-exponentiële factor in de reactiesnelheidscoëfficiënt van de reactie C2H2 + OH →

CH2CO +H naar 1.7 x 1012 cm3mol-1s-1 in het gebied 1760-1850 K leidt tot zeer goede

overeenkomst tussen de berekende en gemeten resultaten.

In hoofdstuk 6 wordt de vorming en afbraak van HCN in brandstofrijke

methaan/lucht vlammen bij atmosferische druk besproken. Hiertoe zijn axiale HCN-

en temperatuur- profielen gemeten voor de equivalentieverhoudingen ϕ = 1.3, 1.4 en

1.5. Voor de brandstofrijkste vlam (ϕ = 1.5), is een zeer langzame oxidatie van HCN

waargenomen in de postvlam-zone, waarmee “resterend” HCN in de postvlam-gassen

aangetoond is. De HCN-metingen laten zien dat het verhogen van de vlamtemperatuur

bij constante equivalentieverhouding een gering effect heeft op de HCN-

piekmolfractie, maar dat door de verhoogde temperatuur de HCN oxidatie substantieel

versnelt in de postvlam-zone. Berekende HCN profielen op basis van het GRI 3.0

mechanisme laten grote verschillen zien met de meetresultaten. Pogingen om de

berekeningen beter in overeenstemming te brengen met de meetresultaten, door

middel van het veranderen van snelheden van belangrijke formatie en consumptie

reacties zijn onsuccesvol gebleken. Door in plaats van HCN, NCN als primair product

van de reactie CH en N2 te beschouwen, is verbetering tussen berekende en gemeten

HCN molfracties mogelijk. Echter, het gebrek aan informatie over

Page 148: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Samenvatting

143

reactiesnelheidsconstanten van de reacties van NCN bij hoge temperatuur verhindert

het trekken van duidelijke conclusies over dit mechanisme.

Hoofdstuk 7 verlegt de inhoud van Hoofdstukken 5 en 6 en beschrijft het effect

van de toevoeging van waterstof op de formatie en afbraak van HCN en C2H2 in

brandstofrijke methaan/lucht vlammen. De HCN-metingen bij ϕ = 1.4 en 1.5 laten

zien dat het verhogen van de waterstoffractie in het mengsel bij constante

vlamtemperatuur resulteert in een substantiële verlaging in de HCN molfractie. Deze

verlaging in HCN is significant groter dan de verlaging die verwacht wordt door de

verdunning van het methaan als gevolg van waterstoftoevoeging. Dit duidt aan dat de

vormingschemie van HCN wordt beïnvloed door de toevoeging van waterstof. Net als

waargenomen voor pure methaanvlammen (Hoofdstuk 6), voorspellen de

berekeningen op basis van het GRI 3.0 mechanisme bij ϕ = 1.4 substantieel hogere

piekmolfracties van HCN dan gemeten, en is de voorspelde oxidatiesnelheid

langzamer. Daarnaast, in tegenstelling tot de experimentele waarnemingen, wordt

slechts een geringe invloed voorspeld op de berekende piekmolfractie van HCN bij

het toevoegen van waterstof en ook een sterke verlaging in HCN molfractie bij

dalende vlamtemperatuur bij constante waterstoffractie. Goede overeenkomst tussen

de metingen en berekeningen is gevonden voor de vlammen bij ϕ = 1.5. Echter, de

reductie in de berekende HCN-molfractie door toevoeging van waterstof is

hoogstwaarschijnlijk het resultaat van de berekende reductie in vlamtemperauur en

niet het gevolg van een “waterstofeffect”. Voor de equivalentieverhoudingen ϕ = 1.3

en 1.5 zijn geen significante veranderingen waargenomen in de gemeten C2H2-

molfracties als gevolg van toevoeging van waterstof. Dit suggereert dat

waterstoftoevoeging geen significant effect heeft op de chemie verantwoordelijk voor

C2H2 in methane/lucht vlammen. Vervangen van de reactiesnelheidscoëfficiënt van de

reactie C2H2 + OH → CH2CO +H in het GRI 3.0 mechanisme door de waarde

voorgesteld in Hoofdstuk 6 levert voor alle bestudeerde vlammen zeer goede

overeenkomsten tussen de gemeten en berekende C2H2 profielen.

Page 149: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Dankwoord

144

Promoveren is in tegenstelling tot wat veel mensen denken geen individualistisch

karwei, maar een proces dat in samenwerking met vele anderen tot stand komt.

Iedereen die een bijdrage heeft geleverd aan dit proefschrift wil ik bij deze bedanken

en een paar mensen in het bijzonder.

Ten eerste wil ik Prof. Dr. Howard Levinsky van harte bedanken voor het feit

dat ik bij hem mijn promotieonderzoek heb mogen uitvoeren. Tijdens mijn doctoraal

onderzoek werd mij duidelijk dat ik mij verder wilde verdiepen in de

verbrandingstechnologie en dit was mogelijk dankzij het promotieonderzoek dat je

mij aanbood. Verder wil ik je bedanken voor de leuke en leerzame discussies tijdens

de wekelijks terugkerende vakgroep besprekingen en de tips die je me hebt gegeven

tijdens het schrijven van artikelen en het proefschrift.

Mijn co-promotor dr. Anatolia Mokhov wil ik bij deze bedanken voor de intensieve

begeleiding op het gebied van onder andere laser diagnostiek en het schrijven van

artikelen. Tolja, je hebt mij versteld doen staan van je vakkennis en ik ben je zeer

dankbaar dat je deze kennis, waar mogelijk, met mij hebt gedeeld. Verder wil ik je

bedanken voor de prettige samenwerking en de leuke tijd op de Universiteit.

Dit onderzoek is tot stand gekomen dankzij de financiësle steun van het

programma EET (Economie, Ecologie, Technologie), waarvoor dank. De

beoordelingscommissie bestaande uit Prof. dr. ir. R. Baert, Prof. dr. H.C. Moll,

Prof. en dr. ir. Th.H. van der Meer wil ik bedanken voor hun beoordeling van het

proefschrift.

Daarnaast wil Marcel en Edwin van het bedrijf ERMA uitbesteding bijzonder

bedanken voor de constructie van de Rapid Compression Machine. Kees en Ubbel

hebben tijdens mijn promotieonderzoek zeer goed geholpen met onder andere het

leveren van gassen, hiervoor hartelijk dank.

Uiteraard wil ik mijn oud collega’s Alexei, Martijn, Nikolay en Vishal

bedanken voor alle hulp en gezelligheid, hierbij denk ik onder andere aan onze

vrijdagmiddag borrels in de Irish Pub. Especially, I wish to thank Nikolay for his

assistance in performing the autoignition measurements. Martijn, ik denk met veel

plezier terug aan de 4 jaar dat we een kantoortje en lab deelden. Niet alleen de

wetenschappelijke discussies waar ik veel van geleerd heb, maar ook onze befaamde

tafeltenniswedstrijden samen met Alexei waren zeer geslaagd.

Page 150: University of Groningen Experimental study of the ... · However, since the combustion properties of hydrogen differ in many respects from those of natural gas, the allowable fraction

Dankwoord

145

Naast werk was er tijd voor ontspanning, daarvoor wil ik mijn vrienden erg

bedanken. Natuurlijk wil ik mijn familie bedanken en in het bijzonder mijn ouders;

Wim en Anneke voor hun steun en vertrouwen in mij.

Verder wil ik Daniëla speciaal bedanken voor het ontwerpen van de prachtige

omslag van mijn proefschrift.

En natuurlijk wil ik Sandra bedanken voor al haar liefde en energie die ze me heeft

gegeven om door te gaan met mijn promotieonderzoek, met name tijdens de

momenten waarin het onderzoek niet verliep zoals ik dat wilde.