13
Paper # 070RK-0050 Topic: Reaction Kinetics 1 8 th U. S. National Combustion Meeting Organized by the Western States Section of the Combustion Institute and hosted by the University of Utah May 19-22, 2013 A comprehensive experimental and modeling study of iso-pentanol combustion S. Mani Sarathy 1 , Sungwoo Park 1 , Weijing Wang 2 , Peter Veloo 3 , Alexander C. Davis 1 , Casimir Togbe 4 , Bryan Weber 5 , Charles K. Westbrook 6 , Okjoo Park 7 , Guillaume Dayma 4 , Zhaoyu Luo 5 , Matthew A. Oehlschlaeger 2 , Fokion Egolfopoulos 7 , Tianfeng Lu 5 , William J. Pitz 6 , Jackie Sung 5 , Philippe Dagaut 4 1 Clean Combustion Research Center, King Abdullah University of Science and Technology, Kingdom of Saudi Arabia 2 Mechanical, Aerospace, and Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, NY, USA 3 Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, NJ, USA 4 CNRS-INSIS, 1C, Ave de la Recherche Scientifique, Orleans Cedex 2, France 5 Department of Mechanical Engineering, University of Connecticut, Storrs, CN, United States 6 Lawrence Livermore National Laboratory, Livermore, CA, USA 7 Department of Aerospace and Mechanical Engineering, University of Southern California, Los Angeles, CA 90089, USA Abstract Biofuels are considered as potentially attractive alternative fuels that can reduce pollutant emissions. Ethanol is the most commonly used biofuel to power automobiles, but ethanol has several disadvantages such as low energy density, high O/C ratio, and high hygroscopicity. Iso-pentanol is one of next- generation biofuels that can be used as an alternative fuel in combustion engines because of higher energy density and lower hygroscopicity compared to ethanol. In the present study, new experimental data for iso-pentanol in shock tube, rapid compression machine, jet stirred reactor, and counterflow diffusion flame are presented. A detailed chemical kinetic model for iso-pentanol oxidation was developed including high- and low-temperature chemistry for a better understanding the combustion characteristics of higher alcohols. First, bond dissociation energies were calculated using ab initio methods. The proposed rate constants were based on a previously presented model for butanol isomers and n-pentanol. The model was validated against new and existing experimental data in shock tubes, rapid compression machines, jet stirred reactors, premixed flames, and non-premixed flames. Shock tube ignition delay times were measured for iso-pentanol/air mixtures at equivalence ratios of 0.5, 1.0, and 2.0, at temperatures ranging from 790 to 1252 K, and at nominal pressures of 40 and 60 bar. New jet stirred reactor experiments are reported at 5 atm and four different equivalence ratios. Rapid compression machine ignition delay data was obtained at 40 bar, three equivalence ratios, and temperatures below 800 K. The present mechanism shows good agreement with the data obtained from a wide variety of experimental conditions. Premixed laminar flame speeds and non-premixed extinction strain rates were obtained using the counterflow configuration. The method of direct relation graph (DRG) with expert knowledge (DRGX) was employed to eliminate unimportant species and reactions in the detailed iso- pentanol mechanism and then predict non-premixed flame behavior. In additions, reaction path and temperature A-factor sensitivity analyses were conducted for identifying key reactions. Introduction

A comprehensive experimental and modeling study of iso

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

1

8th U. S. National Combustion Meeting Organized by the Western States Section of the Combustion Institute

and hosted by the University of Utah May 19-22, 2013

A comprehensive experimental and modeling study of iso-pentanol

combustion

S. Mani Sarathy1, Sungwoo Park1, Weijing Wang2, Peter Veloo3, Alexander C. Davis1, Casimir Togbe4, Bryan Weber5, Charles K. Westbrook6, Okjoo Park7, Guillaume Dayma4, Zhaoyu Luo5, Matthew A.

Oehlschlaeger2, Fokion Egolfopoulos7, Tianfeng Lu5, William J. Pitz6, Jackie Sung5, Philippe Dagaut4

1 Clean Combustion Research Center, King Abdullah University of Science and Technology, Kingdom of Saudi Arabia

2 Mechanical, Aerospace, and Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, NY, USA 3 Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, NJ, USA

4 CNRS-INSIS, 1C, Ave de la Recherche Scientifique, Orleans Cedex 2, France

5 Department of Mechanical Engineering, University of Connecticut, Storrs, CN, United States

6 Lawrence Livermore National Laboratory, Livermore, CA, USA

7 Department of Aerospace and Mechanical Engineering, University of Southern California, Los Angeles, CA 90089, USA

Abstract Biofuels are considered as potentially attractive alternative fuels that can reduce pollutant emissions. Ethanol is the most commonly used biofuel to power automobiles, but ethanol has several disadvantages such as low energy density, high O/C ratio, and high hygroscopicity. Iso-pentanol is one of next-generation biofuels that can be used as an alternative fuel in combustion engines because of higher energy density and lower hygroscopicity compared to ethanol. In the present study, new experimental data for iso-pentanol in shock tube, rapid compression machine, jet stirred reactor, and counterflow diffusion flame are presented. A detailed chemical kinetic model for iso-pentanol oxidation was developed including high- and low-temperature chemistry for a better understanding the combustion characteristics of higher alcohols. First, bond dissociation energies were calculated using ab initio methods. The proposed rate constants were based on a previously presented model for butanol isomers and n-pentanol. The model was validated against new and existing experimental data in shock tubes, rapid compression machines, jet stirred reactors, premixed flames, and non-premixed flames. Shock tube ignition delay times were measured for iso-pentanol/air mixtures at equivalence ratios of 0.5, 1.0, and 2.0, at temperatures ranging from 790 to 1252 K, and at nominal pressures of 40 and 60 bar. New jet stirred reactor experiments are reported at 5 atm and four different equivalence ratios. Rapid compression machine ignition delay data was obtained at 40 bar, three equivalence ratios, and temperatures below 800 K. The present mechanism shows good agreement with the data obtained from a wide variety of experimental conditions. Premixed laminar flame speeds and non-premixed extinction strain rates were obtained using the counterflow configuration. The method of direct relation graph (DRG) with expert knowledge (DRGX) was employed to eliminate unimportant species and reactions in the detailed iso-pentanol mechanism and then predict non-premixed flame behavior. In additions, reaction path and temperature A-factor sensitivity analyses were conducted for identifying key reactions. Introduction

Page 2: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

2

The interest in alternative fuels and fuel additives has increased for a reduction of emissions in practical engines in recent years. Oxygenated fuels have been considered as alternative fuels in order to reduce NOx and particulate emissions. In addition, the production of oxygenated fuels from renewable sources can balance emissions of the major greenhouse gas (CO2) from combustion devices. However, these oxygenated fuels have to be evaluated in practical applications because some oxygenated fuels need to be regulated prior to application [1]. Ethanol is an attractive alternative bio-based alcohol fuel extender for petroleum fuels. Although ethanol can reduce the dependency upon petroleum fuels and greenhouse gas emissions, disadvantages such as high O/C ratio, high hygroscopicity and low energy density can cause problems. On the other hand, higher alcohols have low hygroscopicity, higher energy density and lower corrosivity, so they can be more blended well with hydrocarbon fuels in fuelling systems. More recently, the interest for using iso-pentanol (3-methyl-1-butanol, isoamyl alcohol) as a bio-derived [2] gasoline substitute has been increased because it has a higher energy density and is less hygroscopic. However, there have been limited studies on the iso-pentanol combustion. Tsujimura et al. studied the fundamental characteristics of iso-pentanol as a fuel for HCCI engines [3] and the results were later modeled by Tsujimura and co-workers using detailed kinetics [6]. Dayma et al. [4] measured the concentrations of reactants, intermediate species and products over a range of equivalence ratios and temperatures in a JSR at 10 atm, and modeled using a detailed chemical kinetic mechanism. Welz and co-workers [5] investigated the low temperature oxidation mechanism of iso-pentanol using photoionization mass spectrometry. Recently, Tsujimura et al. [6] proposed a reaction mechanism for iso-pentanol and the model was validated against ignition delay times measured in a shock-tube at high temperatures and a rapid compression machine at low temperatures. Tang et al. measured the high temperature ignition behavior of iso-pentanol [7] and found good predictions using the model proposed by Dayma [4]. Zhao and coworkers have calculated pressure and temperature dependent rate constants for the thermal decomposition of iso-pentanol [8]. The current study presents a comprehensive data set for iso-pentanol spanning a wide range of temperatures, pressures, equivalence ratios, and experimental configurations. This study also develops a comprehensive chemical kinetic model for iso-pentanol including high- and low-temperature chemistry using rules established for butanol isomers modeling, which have been shown to be suitable for a wide range of conditions. The new model has been validated against recently published experimental data as well as new data presented herein. Chemical kinetic model formulation The detailed chemical kinetic model for iso-pentanol includes both low- and high-temperature chemistry. The proposed model is based on previous modeling study on the combustion of the C4 and C5 alcohols [9,10], and a similar methodology was used in this study to develop a detailed model for iso-pentanol. Only a brief description of the model development process is discussed herein, and readers are referred to [9] and its supplementary material for a more detailed description of the methods employed. The thermodynamic data for iso-pentanol and related radicals were calculated using the THERM program of Ritter and Bozzelli [11]. The detailed chemical kinetic mechanism utilized in this work is based on the hierarchical nature of combustion mechanisms. Therefore, the reaction mechanism of iso-pentanol was developed by adding its primary reactions and related radical reactions to the n-pentanol [10] and iso-butanol [9] reaction mechanisms The chemistry of C5 aldehydes and enols was developed following the methods described in detail for C4 aldehydes and enols in [9,12]. All the validation simulations were conducted in CHEMKIN PRO [13] using the appropriate reactor modules. Quantum chemistry calculations

Page 3: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

3

The reaction rates selected in this study follow directly from previous work on alkanes and alcohols. This approach was initially validated using quantum chemistry calculations to verify the most stable iso-pentanol conformation and to obtain C–H and C–C bond dissociation energies. Consistent with previous studies on n-alcohols, the lowest energy configurations, G+_TG-t and T_G-G+t, have the CCCO in a gauche configuration [10,14,15]. However, iso-pentanol contains a branching methyl group on the γ-carbon that exhibits a strong steric effect on the stability a given configuration. The α-C-H BDE for iso-pentanol and n-pentanol are very similar at 93.6 and 94.0 kcal mol-1, respectively, suggesting that the methyl group has minimal influence on the strength of this bond (Figure 1) [10]. Similarly, the primary C-H (C4-H and C5-H) BDE are 100.4 and 100.5 kcal mol-1, respectively, compared to 100.4 kcal mol-1 in n-pentanol. This is consistent with the observation that the effect of the alcohol group on BDE in n-pentanol is limited up to the γ-C site. However, unlike n-pentanol, the γ-C site in iso-pentanol is a tertiary site with a C-H BDE of 95.2 kcal mol-1, which is only 1.6 kcal mol-1 higher than the α-C-H BDE. This is similar to the analogous site in iso-butanol [9]. For the C-C BDE, similar to n-pentanol, the Cα-Cβ in iso-pentanol has the lowest BDE at 86.0 kcal mol-1, which is nearly identical to the corresponding bond in n-pentanol, 85.9 kcal mol-1. The remaining C-C BDE values are all higher than the Cα-Cβ, though interestingly, the CBS-QB3, G3 and G4 disagree on the extent to which the C3-C4 (and C3-C5), and C2-C3 BDE differ from each other and from Cα-Cβ. According to the CBS-QB3 method, the two bond scissions that lead to the formation of a methyl radical (the C3-C4 and C3-C5) are 0.5 kcal mol-1 lower than the C2-C3, 1.1 kcal mol-1 above Cα-Cβ. According to the G3 data set, the C3-C4 and C3-C5 BDE are 1.5 kcal mol-1 lower than the C2-C3 and only 0.6 kcal mol-1 above the Cα-Cβ. The G4 values place the C3-C4 and C3-C5 BDE slightly closer to the middle, with values 0.7 kcal mol-1 below the C2-C3, and 1.1 kcal above the Cα-Cβ. In each case the scission of the C-C bond leads to the formation of a secondary radical site resulting in lower BDE values than are observed in n-butanol or n-pentanol. As a result scission of these bonds is expected to play a slightly larger role in the overall mechanism of iso-pentanol, and other methylalcohols than it does in n-alcohols like n-butanol or n-pentanol

Figure 1 - Schematic diagram of the bond dissociation energies for iso-pentanol. The numbers in the brackets are G4 values, those in the parentheses are G3 values and those without the brackets are the CBS-QB3 values. Blue values are C–H and black are C–C BDEs (kcal mol–1 at 298 K). Reaction classes and rate rules The major classes of elementary reactions considered for the oxidation of iso-pentanol are as follows: High-temperature reaction classes 1. Unimolecular fuel decomposition 2. H-atom abstraction from the fuel 3. Fuel radical decomposition 4. Fuel radical isomerization 5. H-atom abstraction reactions from enols (i.e., unsaturated alcohols)

Page 4: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

4

6. Enol-Keto tautomerizations and isomerizations catalyzed by H, HO2, and formic acid 7. Addition of H atoms to enols 8. Enol radical decomposition 9. Unimolecular decomposition of enols 10. Reaction of O2 with alpha-hydroxypentyl radicals to directly form an aldehyde + HO2 Low-temperature reaction classes (R refers to an iso-pentanol radical such as (CH3)2C.CH2CH2OH and QOOH refers to a iso-pentyl-hydroperoxide radical such as (CH3)2C(OOH)CH2CH.OH) 11. Addition of O2 to fuel radicals (R + O2 = ROO) 12. R + ROO = RO + RO 13. R + HO2 = RO + OH 14. R + CH3O2 = RO + CH3O 15. ROO radical isomerization (ROO = QOOH) including Waddington type reaction mechanism 16. Concerted eliminations (ROO = enol + HO2) 17. ROO + HO2 = ROOH + OH 18. ROO + H2O2 = ROOH + HO2 19. ROO + CH3O2 = RO + CH3O + O2 20. ROO + ROO = RO + RO + O2 21. ROOH = RO + OH 22. RO decomposition 23. Formation epoxy alcohols via cyclization 24. QOOH = enol + HO2 (radical site beta to OOH group) 25. QOOH = alkene/enol + carbonyl + OH (radical site gamma to OOH group) 26. Addition of O2 to QOOH (QOOH + O2 = OOQOOH) 27. Reaction of O2 with alpha-hydroxyiso-pentylhydroperoxide radicals (e.g., CH3CH2CH(OOH)CH2CH.OH + O2) 28. Isomerization of OOQOOH and formation of carbonyl hydroxyalkyl-hydroperoxide species and OH including Waddington type reactions mechanism 29. Decomposition of carbonyl hydroxyalkyl-hydroperoxide species to form oxygenated radical species and OH 30. Cyclic oxygenates reactions with OH and HO2 As mentioned previously, the CBS-QB3 values for the scission of primary C–H bonds are like values for similar bonds in alkanes, while the values for the OH, α, β, and γ sites are consistent with values for n-pentanol, n-butanol and iso-butanol. Similar trends exist for C–O and C–C bonds in the iso-pentanol molecule. Therefore, iso-pentanol displays the alkane-like and alcohol-specific portions of the molecule, which forms the basis for the allocation of reaction rate constants. Transport Properties The transport properties for new iso-pentanol related species are determined as follows. For stable species, this study uses the correlations developed by Tee, Gotoh, and Stewart [16] as first described by Wang and Frenklach [17] for aromatics, and later by Holley and co-workers for hydrocarbons [18], to calculate the Lennard Jones collision diameter and potential well depth using the critical pressure (pC), critical temperature (TC), and acentric factor (ω). The estimation of the acentric factor (ω) is based on Lee-Kesler vapor-pressure relations, which requires the critical pressure (pC), critical temperature (TC), and boiling point (Tb) of the species. Model Reduction The method of direct relation graph (DRG) with expert knowledge (DRGX) [19-21] was employed to eliminate unimportant species and reactions in the detailed iso-pentanol mechanism, so that it could be used to simulate extinction of counterflow diffusion flames. Compared to DRG [22], DRGX allows users

Page 5: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

5

to specify species and heat release specific error tolerances (i.e., the expert knowledge, or x-value), and can retain selected important species and associated reaction pathways with higher-than-default accuracy, offering a flexibility to achieve balanced chemical fidelities among different reaction groups. Model validation studies The proposed model for the iso-pentanol is validated against a wide range of experimental data covering low-temperature and high-temperature oxidation conditions. The following is a list specific validation targets presented in this study: 1. Auto-ignition data for iso-pentanol using a shock tube (ST) at NUI Galway and rapid compression machine (RCM) at the University of Connecticut over a temperature range of 652–1457 K, pressures near 7 atm and 20 atm, and equivalence ratios of 0.5, 1.0 and 2.0 in air [6]. 2. Concentration profiles of iso-pentanol oxidation in a jet-stirred reactor (JSR) at CNRS Orleans operating at 10 atm, constant residence time of 0.7 s, equivalence ratio of 1.0, and temperatures from 530 to 1220 K [4]. 3. New auto-ignition data in RCM at the University of Connecticut at pressure of 40 atm, temperatures range of 651 – 776 K, and equivalence ratios of 0.5, 1.0 and 2.0 in air 4. New concentration profiles data of iso-pentanol oxidation in a jet-stirred reactor (JSR) at CNRS Orleans operating at 5 atm, constant residence time, τ, of 0.35 s over a range of equivalence ratios (0.35 ~ 4) and temperatures from 530 to 1220 K [4]. 5. New counterflow flame performed measurements for laminar flame speed and extinction strain rate. Experimental Methodologies Shock tube ignition delay Measurements of ignition delay times were carried out in a heated high-pressure shock tube at Rensselaer Polytechnic Institute using the reflected-shock technique. The shock tube has been previously described by Wang and Oehlschlaeger [23] and references therein, hence, only details relating to the present study are provided here. Ignition delay times were measured for iso-pentanol/air mixtures at equivalence ratios of 0.5, 1.0, and 2.0, at temperatures ranging from 790 to 1252 K, and at nominal pressures of 40 and 60 bar. Rapid compression machine ignition delay The present Rapid Compression Machine (RCM) at the University of Connecticut has been described extensively elsewhere [24]. The configuration is a pneumatically-driven, hydraulically-stopped single piston arrangement. The temperature and pressure at piston top dead center (TDC) are independently adjustable by adjusting the stroke of the piston, TDC clearance, initial pressure and initial temperature. The conditions in the reaction chamber when the piston has reached TDC are referred to as the compressed conditions, represented by the symbols PC and TC for pressure and temperature respectively. In this study, the ignition delay in the RCM is defined as the difference in time between the end of compression and the maximum of the time derivative of the pressure. Jet stirred reactor experiments The jet stirred reactor (JSR) experimental setup used here has been described earlier [4,25]. The JSR consists of a 4 cm diameter fused silica sphere (42 cm3) equipped with four nozzles of 1 mm i.d.. Steady state experiments were performed at a constant mean residence time (τ= 0.35 s) and pressure (5 atm), with variable equivalence ratio (0.35 ≤φ ≤ 4).The reactants flowed continually in the reactor and the temperature of the gases inside the JSR was increased stepwise. A good repeatability of the measurements and a good carbon balance (typically 100±10%) were observed. Counterflow flame measurements for laminar flame speed and extinction strain rate

Page 6: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

6

Experiments were conducted in steady, laminar, and planar counterflow flames at atmospheric pressure, p=1, and an unburned reactant temperature, Tu=353K. In order to determine

Suo’s the symmetric twin-

flame configuration was used. The diameters of the burners used are 14 mm for flame propagation experiments and 10 mm for extinction limit measurements. The separation distance between the burners was equal to the burner diameter. The axial velocity profile along the system centerline just upstream of the flame is first measured. The maximum absolute value of the velocity gradient along the centerline just upstream of the flame is defined as the local strain rate, K, and the minimum point of the axial velocity profile just upstream of the flame is defined as the reference flame speed, Su,ref . Plotting Su,ref against K,

Suo could be determined by non-linearly extrapolating Su,ref to K = 0 using a computationally-

assisted approach. To determine non-premixed flame extinction strain rates, Kext, a heated fuel/N2 mixture was counterflowed against O2 jet. Pure O2 was chosen as the oxidizer in order to establish flames of small fuel to N2 ratios in the fuel stream. A flame is established first for a near-extinction condition, and the axial velocity profile along the stagnation stream is measured on the fuel/N2 stream to determine the prevailing local strain rate, K. Subsequently, the fuel flow rate is slightly modified to achieve extinction. Thus, the value of K measured before the fuel flow rate reduction is defined as the experimentally determined extinction strain rate, Kext. Kext’s were measured on the side of the fuel-containing stream. Results and discussions Auto-ignition delay measurements in RCMs and STs Tsujimura et al. [6] measured auto-ignition data for iso-pentanol using a shock-tube at NUI Galway and a rapid compression machine at the University of Connecticut over a temperature range of 652–1457 K, pressure range from 7 to 20 atm and equivalence ratio of 0.5, 1.0 and 2.0. The present model is validated against the high-pressure autoignition experimental data by running homogeneous batch reactor simulations. Figure 2 and 3 present the experimental data obtained in a high-pressure shock tube (ST) and rapid compression machine (RCM) along with computed ignition delay times at 7 and 20 atm for equivalence ratios of 0.5, 1.0 and 2.0. Overall, the proposed model for iso-pentanol is in good agreement with the experimental data over the whole temperature region. A temperature sensitivity analysis at the time of ignition was conducted at 20 atm and 800 K for equivalence ratio of 1.0 and sensitive reactions are given in Figure 4. A positive sensitivity coefficient means that increasing the reaction rate constants will increase the reactivity of the system (i.e., decrease ignition delay time). The results reveal that the hydrogen abstractions reactions with OH and HO2 radicals and the branching fractions are most sensitive at high pressure and intermediate to low-temperatures. In addition, the addition of molecular oxygen to fuel radicals (R + O2 = ROO) and the isomerizations of hydroxyalkyl peroxy radicals to produce hydroxyalkyl hydroperoxide (ROO = QOOH) are also sensitive which indicates that fuel radicals produced from the H-abstractions are mainly consumed by low-temperature chain branching process.

Page 7: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

7

Figure 2 - ST and RCM ignition delay times for iso-pentanol at 7 atm compared with model prediction.

Figure 3- ST and RCM ignition delay times for iso-pentanol at 20 atm compared with model prediction.

Figure 4 - Temperature A-factor sensitivities at the time of ignition for iso-pentanol combustion. ST

conditions are 800 K and 20 atm and equivalence ratio 1.0.

0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.41000/T (1/K)

1.0e-05

1.0e-04

1.0e-03

1.0e-02

1.0e-01

1.0e+00

Igni

tion

Del

ay (s

ec)

phi=0.5 ST RCM Const Vol Vol historyphi=1.0 ST RCM Const Vol Vol historyphi=2.0 ST Const Vol

iso-pentanol in air, 7 atm

0.6 0.8 1 1.2 1.4 1.61000/T (1/K)

1.0e-05

1.0e-04

1.0e-03

1.0e-02

1.0e-01

1.0e+00

Igni

tion

Del

ay (s

ec)

phi=0.5 ST RCM Const Vol Vol historyphi=1.0 ST RCM Const Vol Vol historyphi=2.0 ST RCM Const Vol Vol history

iso-pentanol in air, 20 atm

!800$ !600$ !400$ !200$ 0$ 200$ 400$ 600$ 800$

ic5h11oh+oh<=>ic5h10oh!1+h2o$

ic5h10oh!1+o2<=>ic4h9cho+ho2$

ho2+ho2<=>h2o2+o2$

ic5h9oh1!3+ho2<=>ic5h9oh!4ooh!3$

ic5h10oh!4o2<=>ic5h9oh1!3+ho2$

ic4h9cho+ho2<=>ic4h9co+h2o2$

ic5h11oh+ho2<=>ic5h10oh!3+h2o2$

ic5h9oh!4ooh!3o2<=>ic5ohket4!3+oh$

ic5h10oh!4o2<=>ic5h9oh!4ooh!2$

h2o2(+m)<=>oh+oh(+m)$

ic5h11oh+ho2<=>ic5h10oh!1+h2o2$

ic5h10oh!3o2<=>ic5h9oh!3ooh!1$

ic5h10oh!1+o2<=>ic5h10oh!1o2$

ic5h11oh+oh<=>ic5h10oh!4+h2o$

ic5h11oh+oh<=>ic5h10oh!3+h2o$

Page 8: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

8

The main iso-pentanol reaction pathways after 20% fuel consumption at 800 K, 20 atm for equivalence ratio of 1.0 are shown in Figure 5, describing the key low temperature reaction pathways. Fuel is mainly consumed by the H atom abstraction at the α site because iso-pentanol has a weak C–H bond at the α site. Increasing the rate of abstraction by OH from the α site (ic5h10oh-1) increases the ignition delay time because subsequent reaction with O2 leads to the formation of HO2 and iso-pentanal, which is an OH terminating pathway. As mentioned above, fuel radicals commonly add molecular oxygen at the radical site as a first step in the chain branching process. The exception is α-hydroxypentyl radicals producing iso-pentanal by the reaction with molecular oxygen. Hydroxyalkyl peroxy radicals produced from the addition of O2 to fuel radicals are mainly isomerized to hydroxyalkyl hydroperoxide (QOOH) and decomposed to enol species by the concerted elimination of HO2. Around 18% of β hydroxyalkyl peroxy radical are decomposed to produce isobutanal, formaldehyde and OH radical via ROO isomerization and β scission reactions by the Waddington mechanism [26].

Figure 5 - Reaction path analysis for iso-pentanol at 800 K, 20 atm and Φ = 1.0. The reaction fluxes are

given for 20% fuel consumption The model was also validated against new ignition delay data in RCMs and STs at 40 atm and 60 atm, as shown in figures 6 and 7. The model is able to well predict the shock tube ignition delay data at both pressures. However, the model has difficulty predicting RCM ignition delay data at phi=0.5 and 2. New data in RCMs and STs The model was also validated against new ignition delay data in RCMs and STs at 40 atm and 60 atm, as shown in figures 6 and 7. The model is able to well predict the shock tube ignition delay data at both pressure. However, the model has difficulty predicting RCM ignition delay data at phi=0.5 and 2.

OH#

OH# OH# OH# OH# O#

!!49#%#!!

!!!8#%#!!

!!19#%#!!

!!23#%#!!

!!!1#%#!!

O#+#HO2#

!!!!11#%##!!

OH#

OO#

+#O2#!!!!87#%##!!

!!!!97#%##!!!

OH#OO#

+#O2#

!!!!6#%####

OH#OO#

!!!!#94#%##!!!!

+#CH2O#

QOOH#Isopentenol##+#HO2#

O#

+#CH2O#+#OH#

!!!!81#%##!!!

QOOH#Isopentenol##+#HO2#

!!!!87#%##!!!

QOOH#

!!!!9#%##!!!

Isopentenol##+#HO2#OOH#

+#CH2O#

!!!!6#%##!!!

+#O2# +#O2#!!!!98#%##!!`!

!!!!97#%##!!!

OH#

OO#

+#O2#

QOOH# Isopentenol##+#HO2#

!!!60#%##!!`!

!!!!18#%##!!!

!!!!19#%##!!!

!!!!11#%##!!!

!!!!85#%##!!!

!!!!12#%##!!!

+#O2#

Page 9: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

9

Figure 6 - ST and RCM ignition delay times for iso-pentanol at 40 atm compared with model prediction.

Figure 7 - ST ignition delay times for iso-pentanol at 60 atm compared with model prediction.

Jet-stirred reactor measurements Dayma et al. [4] measured concentration profiles of iso-pentanol oxidation in a jet-stirred reactor (JSR) at 10 atm, constant residence time, τ, of 0.7 s over a range of equivalence ratios (0.35 ~ 4) and temperatures from 530 to 1220 K. New data was also acquired in the JSR at 5 atm. The present model is validated against JSR data for an equivalence ratio of 1.0 at both pressures. Figures 9 and 10 shows comparisons of mole fractions of reactants (iso-pentanol and O2), major products (CO, CO2, H2 and H2O) and intermediate species (methane, ethylene, propene, isobutene, formaldehyde, acetaldehyde and iso-pentanal). The predictions show good agreement with experimental data. However, the simulation results reveal that the fuel consumption occurs at slightly lower temperature than measurement. As can be seen in Figures 9 and 10, iso-pentanal (iC4H9CHO) is over predicted by approximately a factor of 2 at stoichiometric mixture conditions. According to the iso-pentanol reaction path analysis at 800 K

0.8 1 1.2 1.4 1.61000/T (1/K)

1.0e-06

1.0e-04

1.0e-02

1.0e+00

Igni

tion

Del

ay (s

ec)

phi=0.5 ST RCM Const Vol Vol historyphi=1.0 ST RCM Const Vol Vol historyphi=2.0 ST RCM Const Vol Vol history

iso-pentanol in air, 40 atm

0.8 1 1.2 1.4 1.61000/T (1/K)

1.0e-05

1.0e-04

1.0e-03

1.0e-02

1.0e-01

1.0e+00

Igni

tion

Del

ay (s

ec)

phi=0.5

phi=1.0

iso-pentanol in air, 60 atm

Page 10: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

10

and 20 atm, iso-pentanal is a major intermediate species, mainly produced by the addition of molecular oxygen to α-hydroxypentyl radicals.

Figure 8 - iso-Pentanol oxidation in a JSR at 10 atm, t = 0.7 s and phi=1.0. The initial fuel mole fraction

was 0.1%. Experimental data (symbols) are compared to calculations (lines).

Figure 9 - iso-Pentanol oxidation in a JSR at 5 atm, t = 0.35 s and phi=1.0. The initial fuel mole fraction

was 0.1%. Experimental data (symbols) are compared to calculations (lines). Counterflow flame measurements of laminar flame speed and extinction strain rate

0.0e+00

5.0e-04

1.0e-03

1.5e-03iC5H11OH

H2

C2H4

C3H6

0.0e+00

5.0e-03

1.0e-02

1.5e-02O2

CO

CO2

H2O

700 800 900 1000 1100Temperature [K]

0.0e+00

1.0e-04

2.0e-04

3.0e-04

4.0e-04

Mol

e fra

ctio

n [-]

CH4

CH2O

C2H2 (X5)

iC4H8

700 800 900 1000 1100 12000.0e+00

5.0e-05

1.0e-04

CH3CHO

C2H3CHO

iC4H9CHO

iC3H7CHO (X5)

0.0e+00

5.0e-04

1.0e-03

1.5e-03

2.0e-03iC5H11OH

H2

C2H4

C3H6

0.0e+00

2.0e-03

4.0e-03

6.0e-03

8.0e-03O2 (/2)

CO

CO2

H2O

700 800 900 1000 1100Temperature [K]

0.0e+00

2.0e-04

4.0e-04

6.0e-04Mol

e fra

ctio

n [-]

CH4

CH2O

C2H2 (X5)

iC4H8

700 800 900 1000 1100 12000.0e+00

1.0e-04

2.0e-04

CH3CHO

C2H3CHO

iC4H9CHO (X2)

iC3H7CHO (X10)

Page 11: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

11

New experimental data was acquired for laminar flame speeds and extinction strain rates, as shown in Figures 10 and 11, respectively. The skeletal model generated using DRGX was used to simulate these one-dimensional flames. The model under predicts laminar flame speeds of iso-pentanol at equivalence ratios below 1.2, while well predicting them at richer conditions. It is interesting to note that the iso-pentanol model predicts flame speeds similar to measured data for iso-butanol at the same unburnt gas temperature of 353 K [27]. Previous model predictions of iso-butanol [9] show similar agreement with this experimental data. Therefore, the models for iso-pentanol and iso-butanol suggest that the two fuels have similar flame speeds, which is due to the fact that both models are developed using similar reaction classes and rate rules. This consistency of flame speeds for fuels of similar structure but different chain length has been shown experimentally and computationally for n-alkanes [28]. The model also well predicts extinction strain rates at lower fuel mole fractions, but over predicts them at higher fuel mole fractions, as shown in Figure 11.

Figure 10 – Laminar flame speeds at unburnt gas temperature of 353 K. Lines show model predictions for iso-

pentanol. Experimental data for iso-butanol data [27] is also presented.

Figure 11 – Extinction strain rates for iso-pentanol counterflow flames

0

10

20

30

40

50

0.6 0.8 1 1.2 1.4 1.6

Lam

inar

Fla

me

Velo

city

(cm

/s)

Equivalence Ratio

iso-pentanol 353 K USC

iso-pentanol model

iso-butanol 353 K Princeton

200

300

400

500

600

700

800

900

0.028 0.03 0.032 0.034 0.036

Ext

inct

ion

Str

ain

Rat

e, K

ext,

s-1

Fuel Mass Fraction (Xf)

experiments

simulation

Page 12: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

12

Conclusions In summary, a new model for iso-pentanol was validated against new and existing experimental data in shock tubes, rapid compression machines, jet stirred reactors, premixed flames, and non-premixed flames. Shock tube ignition delay times were measured for iso-pentanol/air mixtures at equivalence ratios of 0.5, 1.0, and 2.0, at temperatures ranging from 790 to 1252 K, and at nominal pressures of 40 and 60 bar. New jet stirred reactor experiments are reported at 5 atm and four different equivalence ratios. Rapid compression machine ignition delay data was obtained at 40 bar, three equivalence ratios, and temperatures below 800 K. The present mechanism shows good agreement with the data obtained from a wide variety of experimental conditions. Premixed laminar flame speeds and non-premixed extinction strain rates were obtained using the counterflow configuration. The method of directed relation graph (DRG) with expert knowledge (DRGX) was employed to eliminate unimportant species and reactions in the detailed iso-pentanol mechanism and then predict non-premixed flame behavior. In additions, reaction path and temperature A-factor sensitivity analyses were conducted for identifying key reactions. References [1] A. Agarwal, Progress in Energy and Combustion Science 33 (2007) 233. [2] P. Peralta-Yahya, J. Keasling, Biotechnology Journal 5 (2010) 147. [3] T. Tsujimura, W. Pitz, Y. Yang, J.E. Dec, International Journal of SAE. SAE 11ICE-0303/2011-

24-0023 (2011). [4] G. Dayma, C. Togbe, P. Dagaut, Energy Fuels (2011). [5] O. Welz, J. Zádor, J.D. Savee, M.Y. Ng, G. Meloni, R.X. Fernandes, L. Sheps, B.A. Simmons,

T.S. Lee, D.L. Osborn, C.A. Taatjes, Phys Chem Chem Phys 14 (2012) 3112. [6] T. Tsujimura, W.J. Pitz, F. Gillespie, H.J. Curran, B.W. Weber, Y. Zhang, C.-J. Sung, Energy

Fuels (2012). [7] C. Tang, L. Wei, X. Man, J. Zhang, Z. Huang, C.K. Law, Combustion and Flame (2013) 1. [8] L. Zhao, L. Ye, F. Zhang, L. Zhang, J. Phys. Chem. A (2012). [9] S.M. Sarathy, S. Vranckx, K. Yasunaga, M. Mehl, P. Osswald, W.K. Metcalfe, C.K. Westbrook,

W.J. Pitz, K. Kohse-Hoinghaus, R.X. Fernandes, H.J. Curran, Combustion and Flame 159 (2012) 2028.

[10] K.A. Heufer, S.M. Sarathy, H.J. Curran, A.C. Davis, C.K. Westbrook, W.J. Pitz, Energy Fuels (2012).

[11] E. Ritter, J. Bozzelli, Int. J. Chem. Kinet. 23 (1991) 767. [12] K. Yasunaga, T. Mikajiri, S.M. Sarathy, T. Koike, F. Gillespie, T. Nagy, J.M. Simmie, H.J.

Curran, Combustion and Flame (2012). [13] Reaction Design, CHEMKIN PRO 15112 (2012). [14] J. Moc, J.M. Simmie, H.J. Curran, Journal of Molecular Structure 928 (2009) 149. [15] X. Xu, E. Papajak, J. Zheng, D.G. Truhlar, Phys Chem Chem Phys 14 (2012) 4204. [16] L. Tee, S. Gotoh, W. Stewart, Ind Eng Chem Fund 5 (1966) 356. [17] H. Wang, M. Frenklach, Combustion and Flame 96 (1994) 163. [18] A. Holley, X. You, E. Dames, H. Wang, F. Egolfopoulos, Proceedings of the Combustion

Institute 32 (2009) 1157. [19] W. Liu, R. Sivaramakrishnan, M.J. Davis, S. Som, D.E. Longman, T.F. Lu, Proceedings of the

Combustion Institute (2013). [20] S.M. Sarathy, U. Niemann, C. Yeung, R. Gehmlich, C.K. Westbrook, M. Plomer, Z. Luo, M.

Mehl, W.J. Pitz, K. Seshadri, M.J. Thomson, T. Lu, Proceedings of the Combustion Institute (2012) 1.

[21] T. Lu, M. Plomer, Z. Luo, S.M. Sarathy, W.J. Pitz, S. Som, D.E. Longman, 2011 7th US National Combustion Meeting (2011) 1.

[22] T. Lu, C. Law, Proceedings of the Combustion Institute 30 (2005) 1333. [23] H. Wang, M.A. Oehlschlaeger, Fuel (2012) 1.

Page 13: A comprehensive experimental and modeling study of iso

Paper # 070RK-0050 Topic: Reaction Kinetics

13

[24] G. Mittal, C.-J. Sung, Combustion Science and Technology 179 (2007) 497. [25] P. Dagaut, M. Cathonnet, J. Rouan, R. Foulatier, A. Quilgars, J. Boettner, F. Gaillard, H. James,

Journal of Physics E: Scientific Instruments 19 (1986) 207. [26] D.J.M. Ray, W.D. Waddington, Combustion and Flame 20 (1973) 327. [27] W. Liu, A. Kelley, C. Law, Proceedings of the Combustion Institute (2010). [28] C. Ji, E. Dames, Y.L. Wang, H. Wang, F.N. Egolfopoulos, Combustion and Flame (2009) 1.