You are on page 1of 12
Combustion and Fame 158 (2011) 1444-1455, Contents lists available at ScienceDirect Combustion and Flame .com/locate/combustflame ELSEVIER journal homepage: www.else Role of peroxy chemistry in the high-pressure ignition of n-butanol - Experiments and detailed kinetic modelling S. Vranckx®*, K.A. Heufer®, C. Lee*, H. Olivier*, L. Schill”, W.A. Kopp”, K. Leonhard, CA. Taatjes‘, RX. Fernandes" shock Wow Laboratory RTH Aachen Univers, Telerraben 55, 52056 Aachen, Germany = recncl Thermodynamics (117) HVT Aachen Univer, Tenplergaten 5, 52056 Aachen, Germany “Combuton Besurch Fact, MIS 9055, Sie Natona Laboratori, Livermore, CA ASS1-069, USA ARTICLE INFO ABSTRACT Received 9 September 2010 Received in revied form 14 October 2010 ‘eeepted 27 December 2010 ‘alae cline Febroay 2011 Despite considerable interest in butanol as a potential biofuel candidate, its ignition behaviour at elevated pressures still remains largely unexplored. The present study investigates the oxidation of ‘butanol in air at pressures near 80 ba. Ignition delays were determined experimentally in the temper Sure range of 795-1200 K between 61 and 92 bar. The time of ignition was determined by recording pressure and CH-emission time histories throughout the course ofthe experiments. The results display the first evidence ofthe influence of negative temperature coefficient (NTC) behaviour which was not observed in earlier ignition studies. The high-pressure measurements show that NTC behaviour is enhanced as pressures are increased. The experimental results were modelled using an improved chemical kinetic mechanism which includes a simplified sub-mechanism for buty-peroxy formation Keywords ‘Burana oxidation high pressure kinetics Peony chemistry Tenten delays and jsomerisation reactions currently incompletely accounted for in n-butanol kinetic models. The Shock abe Getaled mechanism validated with the high-pressure ignition results for realistic engine in-cylinder ‘conditions can have significa impact on future advance low-temperature combustion engines. '© 2011 The Combustion Institute. Published by Elsevier Inc. llrightsreserved, 1. Introduction A similar approach in biofuel research is adopted by the Biofuels are currently gaining increased importance worldwide ina continuously changing fuel stream and have been projected as attractive alternatives to conventional fossil fuels in the tanspor- tation sector. They provide a means to address rising energy de- ‘mand, climate change and elevated CO, emissions as they hold a promise of renewability and substantial reduction in greenhouse ‘gases [1] Current biofuel-related literature is vast and surveys di- Verse aspects both on the production [2-6] as well as on the com- bustion side [7-15]. Many new potential biofuel candidates are proposed in the literature [1,4,16-21] which could be good substi- tutes to fosil-based fuels for both spark ignition (St) and compres- sion ignition (CI) based engines in the near future. New biofuel energy research centres are being developed both in Europe [22,23] as well as in the USA [24] with a primary mission to ad- vvance the development ofthe next generation of biofuels, focusing, on the frontiers between basic and applied science and on bioener- ‘ay applications and bridging knowledge gaps by a diverse scientific approach. ‘malades: ernest sachende (RX Fernandes) (010.2180 se ron mat recently founded interdisciplinary research institution called “Tailor Made Fuels from Biomass” (TMFB) cluster of excellence [23] at RWTH Aachen University, which focuses on novel synthetic pathways in biofuel! production and their performance in future ow temperature combustion technologies in order to assess their potential and challenges in advanced engines. The current work is one of the contributions from the cluster activities and is focussed towatds the investigation of ignition properties and ‘combustion characteristics of novel biofuels, especially related to future low-temperature combustion technologies. ‘One prominent example of a low-temperature engine concept is Homogenous Charge Compression Ignition (HCCI) which is re- garded as a promising advanced engine technology capable of simultaneously decreasing pollutant formation and increasing en- ‘ergy efficiency. However, a critical issue for HCCI is that ignition is predominantly controlled by chemical kinetics, for which accurate fate data for crucial reactions are decisively needed, Furthermore, the operating pressures in advanced, highly boosted engine con~ cepts can easily reach 100 bar or higher. These pressures present substantial challenges to experimental kineticists to perform high-pressure lab-scale experiments that mimic real engine condi- tions and provide validation tools for detailed chemical kinetic model development. A major challenge for kinetic modelling is ‘© 2011 The Combustion Ist. Publishe by Ever In. A rights reserved S Vrant eta Combustion and Fame 158 (2011) 1644-1455 5 pressure dependence of the branching fractions for the most sen: tive reactions. The high-pressure limit may not be reached even at engine conditions for many reactions [25,26], and high-pressure branching fractions are often unknown, Enhanced pressure effects and bath gas influence can introduce even moze complexities [27] to the currently complex problem, particularly as detailed models generally assume the high-pressure limit, which is very rarely termined experimentally. Therefore, for a detailed understanding of the reaction mecha- nisms influencing the combustion phasing under “real engine” conditions, it still r¢mains necessary to investigate the combustion characteristics of new fuels under wider temperature and pressure ranges. n-Butanol is a widely investigated fuel because of its diverse production pathways [2.45,28-32] and recent studies suggest that it is one of the few alcohol fuels that can be produced from a renewable source in a high yield which could make ita very good substitute to gasoline [1,4]. Due to its high energy density, low hhygroscopicity, and low volatility, n-butanol has a closer resem- blance than ethanol to current conventional fuels such as gasoline nd diese! [33]. This would mean that it could require minimal or rho modifications both to the engine design as well as to fuel deliv ery infrastructure [1,34-26} ‘On the combustion side, n-butanol is already a widely invest gated fuel and determining the combustion characteristics espe- Cally its ignition behaviour isin the focus. The main goal here is 10 develop detailed kinetic models capable in predicting the iar tion behaviour as well as pollutant formation beyond current experimental limitations. Previous experimental studies for n- butanol range from flames [34,35,37~40], static reactors (41), low reactors [42]. jet-stirred reactors [38.43], shock tubes [44-47] and engines [48-53]. Nevertheless, its ignition behaviour at elevated pressures had not been investigated before our recent measurements up 10 42 bar (47) Moss et al. [45] measured the ignition delays for the four iso ‘mers of butanol in shock tubes. They observed different ignition behaviour forall the four isomers at high temperatures for stoic fometric mixtures in the temperature range from 1200 to 1800 K and pressures ranging, from 1 to 4 bar. Reaction-flux and sensitiv- ity analysis were performed to illustrate the importance of classes of consumption reactions, among which dehydration, unimolecu- Jar decomposition and H-atom abstraction were found to be most important. The authors report a detailed kinetic model to describe their own experimental data and good agreement was reported for the relatively low pressure data (below 3.5 bar). Larger dis- agreement (factor of 3 underestimation in ignition delays) with their kinetic model is observed at higher pressures (3.5~4.3 bar) {or more diluted mixtures which the authors speculated was ‘due to falloff effects that were neglected in their mechanism. Fu thermore, the reported deviations in experiments andl madel are ‘prominent only at the lower end of temperatures and at the high- er end of pressures, These studies were followed by another ir tion study by Black et al. [44], also reporting experiments behind reflected shock waves over a temperature range from 1100 to 1800 K Tor pressures up to 8 bar. A detailed chemical kinetic mod- el for bio-butanol was developed and the model was validated against their shock tube ignition delay data and also against jet- stirred reactor experiments from a separate study [43]. Reason- able agreement was reported between experiments and detailed kinetic modelling. It is very important to note that both the above-mentioned studies report discrepancies between experiments and model at _modest pressures close to 8 atm. Despite the fact that engine oper~ ating pressures are substantially higher, there is still a lack of experimental studies at these pressures, This is the main motivation forthe current study which focuses, ‘on the intermediate-temperature regime and very high pressures, precisely where current kinetic models often fail to capture the ignition behaviour of novel fuels. This work extends and clarifies ‘our preliminary experiments [47] of ignition delays of n-butanol oxidation around 40 bar in a similar temperature regime (770- 1250 K) where non-linear dependence of ignition delays on tem- perature were ist observed but for which the exact reasons could not be confirmed, 2. Experimental 2.1 Experimental facility Experiments have been performed using a high-pressure stain- Jess-steel shock tube [54]. The inner diameter ofthis shock tube is 140mm and the shock tube has a total length of approximately 15m, The maximum working pressure of the facility is 1000 bar. The large inner diameter in combination with the high working, pressure makes the Aachen shock-tube facility unique for investi gating ignition delay times for realistic engine applications. The ‘1m long driven section is separated from the driver section by -double-diaphragm chamber housing two stainless-steel dia- pphragms each with a thickness of 3 mm. For the current experi iments the facility is resistively heated up to 110°C to keep the fuel vaporised and avoid condensation prior to the experiment. Stoichiometric gas mixtures in synthetic air (3.3% n-butanol, 19.8% Os, 76.9% No) have been directly prepared in the shock tube using a special home-made fuel injector and delivery system and care has been taken to ensure homogenous mixing of fuel (n-butanol > 98%, Sigma Aldrich) and synthetic air (Westfalen, 79.5% Nz, 20.5% 0,). The pressure is monitored to assure complete vvaporisation of the fuel injected in the previously evacuated tube. ‘The homogeneity of the mixture after its subsequent filling with air is determined by performing test experiments for several mixing. times (from few minutes to several hours) under same conditions Experiments were cartied out in the regime where there was 10 change of ignition delay for increased mixing time. For pressure measurement and the determination of the shock speed, five piezoclectric gauges have been used, which have been shielded with a thin layer of silicone rubber. The silicone rubber coating is essential in damping the heat flux transfer from the hhot gas to the pressure probe. In absence of such a coating, the heat flux would lead to a negative pressure-gradient signal which might ‘obscure the effect of increasing pressure gradient due to shock attenuation (see Fig. 1), and result in misinterpretation ofthe scale of facility-related effects, In out facility, one of the pressure probes has been mounted on, the end wall and the other probes are located at different positions fon the sidewall (10 mm, $10 mm, 1010 mm and 2510 mm from the end wall). The ignition delay time has been determined with the pressure gauge at the end wall In addition to the pressure ‘measurements, side-wall optical access located at 10 mm from the end wall has been used to record CH emissions at 431.5 nm selected by a narrow-band-pass filter using a photo multiplier tube (PMN), All measurements have been recorded with a sampling rate of 2.5 MHz 2.2. Interpretation of shock tube ignition delay data For determining the conditions behind the reflected shock, shock velocities are measured with the pressure probes inthe side wall at different positions. By extrapolating the linear fit from these measurements, the shock velocity at the end wall is deter- ‘mined, which is used in combination with the initial conditions 16 |S ance a Combustion and lame 158 (2011) 1404-1455 ‘of the fuel-air mixture as input parameters fr the shock property calculation program KASIMIR [54]. The deviation between the pressure from the shock tube simulation and the measured pres- sure from experiments is about 2%. ‘The ignition delay time determined both from the pressure pro- file as well s the CH emission history is defined as the time differ- ‘ence between the arrival of the incident shack atthe end wall and the prompt rise in pressure, simultaneous to the strongest increase in the emission intensity (see Fig. 1). For most cases the difference between the ignition times determined from the pressure signal and the emission signal is less than about 2% (e.g. in Fig. 1 itis 2.2%), For high temperatures (>1150 K) and short ignition delays (005 ms), geater deviation up to 9% has been observed, but this, is stil less than the scatter obtained from the experimental data points (see Fig. 2), ‘One important aspect for interpretation of shock-tube ignition delay data is the influence of the so called “facility effects" and pre-ignition. Facility effects, eg. positive pressure and temperature ‘gradients behind the reflected shock resulting from shock attenua- tion, have been observed and described extensively in previous studies {55-58}. The shock attenuation and the resulting pressure and temperature increase depend on the inner diameter of the Shock tube, the shock Mach number and the initial pressure [55] Especially for larger inner tube diameters, shock attenuation and the resulting pressure and temperature gradients are low. Further- ‘more, shock attenuation is reduced for smaller shock Mach num bers and higher initial pressures in the driven section. In addition to shock attenuation causing pressure and temperature gradients, inhomogenous ignition might also occur which can cause local “ignition kernels” resulting in even a further increase in pressure and temperature behind the reflected shock even before the main ‘ignition [59,60] event occurs. These ignition kernels are considered tobe caused by local inhomogeneities or perturbances. One indica- tion of the appearance of pre-ignition is a pressure and emission signal rise before main ignition that can be clearly identified by using Schlieren imaging techniques. Although such experiments are currently underway in our laboratory as a separate study, this is beyond the scope of the present paper. ‘Without the influence of chemistry, both effects, namely shock attenuation and pre-ignition, lead to higher pressures and tem- peratures than they would appear for a constant (U, V) reactor behind the reflected shock, consequently resulting in shorter ign tion delay times. To account for this in kinetic simulations, the “0100 01 02 03 04 05 {ms} ‘ig. Typlcal pressureandCH-emksin tne rfl aitiondeay measurement 00077 (14) Fig. 2. Pressure dependence of ignition delay result for stoichiometric n-butanal- Ar matures. The dashed ies show high-temperacureArents behav program CHEMSHOCK [61] has been developed by the group in ‘Stanford, based on the assumption of an isentropic compression behind the reflected shock. Experiments with H,/O,/Ar mixtures ‘confirm the validity of this approach [57], A recent overview of different studies, describing facility effects, pre-ignition and the Interpretation of shock tube ignition delay data, is presented in [62] and will not be discussed in more detail here. However, it is important to note that these effects can have a large influence ‘on obtained ignition data and need to be considered very care- fully in data. analysis and interpretation. The procedure for detailed data analysis in the current work will be discussed in the following sub-sections. 2.3. Experimental results Inthe present study stoichiometric n-butanol/air mixtures have been investigated. Pressures have been varied from 61 bar to ‘92 bar for temperatures from 795 K to 1200 K. The main focus of the present study lies in the characterisation of low-temperature ‘combustion behaviour of n-butanol. Higher-temperature combus- tion studies are reported elsewhere [44,45,47], and some of this data is used here for model validation 45 —utanol-ar] | 35 30 25 Pip, ms) Fig. 1. Nomaied pressure profes fr stoehemrem bulana ignition delays S Vranl eta Combustion and Fame 155 (2011) 1444-1455 uw Figure 3 depicts typical pressure-time profiles normalised to the pressure behind the reflected shock, py Due to shock attenua- tion a pressure gradient is observed in all experiments, which is on average did (Pipa) ~ 2.5%/ms. For comparison, experiments with air as driven gas have also been performed and a typical pressure profile for such experiments is depicted in the same plot. Under similar conditions behind the reflected shock, an identical pressure gradient is clearly observed, indicating that this gradient is not strongly influenced by chemistry. Furthermore, this experiment shows that the available measuring time, which is defined as the time of constant pressure gradient, amounts to more than 4 ms Following the assumption of an isentropic compression behind the reflected shock [55,56], the pressure gradient behind the re- flected shock corresponds to a temperature gradient of dat (T/ To) =(? ~ I)fy dldt(plpo). This simple relation between the pres- sure and temperature gradient has been verified in experiments of Petersen and Hanson [56]. For the investigated conditions in the present study the isentropic exponent is determined by the KASIMIR program and varies between 1.25 and 1,28 resulting in a temperature gradient of d/dt(T/70)=05-0.55%/ms. Thus, for the longest measured ignition delay time of 2.9 ms the maxirnumn change in temperature is about 13 K Besides the effect of shock attenuation, a pressure rise shortly before the main ignition is possibly due to homogenous exothermi city which is typical for hydrocarbon fuels, Predicted pressure time profiles depicted in Fig. 3 justifies the same. Additionally, such elfects could be due to localized pre-ignition. According to the theory and based on our assumption of an isentropic compression bbchind the reflected shock this would also cause an increase in temperature. Figure 2 shows the experimental results of the present study in, ‘comparison to a previous study performed in the same facility [47]. ‘The figure shows results both as experimentally obtained and as correlated to the average reference pressure using the following formula: se ose = Coa (2) wo ‘The correlation formula, which is derived from the Arshenius equation, is commonly adopted by the shock tube community t0 ‘compare experimental results from other facilities by correlating. the data to a reference pressure. The term “exp” and “cor” in the formula indicate experimentally measured and correlated values respectively. The exponential term °x" refers to a correlation parameter which is determined from the best fit to the obtained experimental resulls. A correlation parameter of x= -0.9 has been ‘obtained in this study which best represents our experimental data ‘over the limited temperature regime investigated. In general, itis ‘inaccurate to apply such a simple correlation over a very wide range of temperature and pressure conditions. This correlation parameter isin close agreement to the factor of 0.7 proposed in an earlier study [47] and is best suited for higher temperatures (1000 K). Our experimental data is summarised in Table 1 for all experiments of this study. For temperatures below 1000 K, the experimental results devi ate from a linear Arrhenius type behaviour. There are several pos- sible reasons for such deviations. Fist, facilty-related effects and pre-ignition can cause shorter ignition delay time, especially at the longer residence times common at lower temperatures. An- fther reason which cannot be ruled out is that a different mecha- nism could be operating in the transition to the intermediate and ower temperature regimes at very high pressures, leading 10 shorter ignition delays. Low-temperature oxidation chemistry ini tiated by the peroxy species also found responsible for the negative temperature coefficient (NTC) has been discussed extensively in the literature for primary reference fuels such as n-heptane. How- ever, in all previous experimental studies for butanol [3841-53] such behaviour has been absent. The present result raise the ques- tion whether the observed results at higher pressures simply arise from facility-induced effects or whether the discrepancy is indeed ‘due to the peroxy chemistry ~ which until now has not been ob- served fo play an important role ~ or whether it could be a combi: nation of the two effects, which would make the data interpretation more difficult. These aspects will be discussed sepa- rately in the following sections of the paper “To separately and reliably account for the facility effects, the absolute change in pressure and temperature behind the reflected shock needs to be measured accurately. As already indicated in Fig. 3 fora given pressure gradient the absolute change of pressure and temperature becomes larger for longer residence times. Be- ‘ause of the large inner diameter and the longer driven section of the shock tube used in the current study, the increase in pres- sure and temperature behind the reflected shock due to shock attenuation remains quite low, and within the maximum mea- ‘ured ignition delay time of about 3 ms this is expected to have a negligible effect on the obtained ignition delay results. However, for longer residence times (>3 ms), these effects could stil have a significant influence. In order to isolate these effects from the fuel chemistry and kinetics, the measured pressure gradients must be included in the kinetic simulations. For chatacterisation of the influence of the facility and pre- ignition effects without any kinetic modelling, “worst-case” con- siderations have been made here assuming solely faclity-induced effects. These “worst-case” assumptions would be a measure for the largest uncertainty in the obtained ignition results due to facility and pre-ignition effects. Therefore, the largest error for the ignition results would appear if the measured ignition delay time does not directly correspond to the conditions behind the ‘Tablet iperimentalconons and ignton dey ines, absalute change of pressure and temperature due wo shock atonal (SA) and preigalon (0. pba me) etm) ae ans) asa ‘ap (a) SAP ar SAT a 856 oe on 03 a1 02 uate |S ance a Combustion and lame 158 (2011) 1404-1455 reflected shock but to the conditions immediately before ignition. Although the pressure rise before the main ignition may also result from chemistry the pressure at gy is used for this worst case con- sideration (see Fig. 4), The temperature is again determined by assuming an isentropic compression behind the reflected shock. ‘The possible shift of the experimental data referred to the condi tions directly before the main ignition is plotted in Fig. 4 (as open Circles). Due to increasing temperature before main ignition the ‘worst case data are shifted towards higher temperatures. Further~ ‘more, the correlated ignition delay time also changes due toa dif ferent pressure using the simple Arrhenius dependency (Eq. (1) ‘A comparison to a kinetic simulation using a previously pub- lished n-butanol kinetic model for high temperatures [44] is shown inFig. 4.In Fig. Sa further comparison is depicted with al available literature models extended to the intermediate temperatures be- ‘yond the regime for which most of the kinetic models have been Validated. We highlight this extrapolation to emphasise the imita- tions in adopting models beyond their validation regime. Itis often difficult to judge the validity regime ofa specific kinetic motel un- less itis very explicitly reported. Moreover, it is also difficult to make an independent « priori estimate of the exact validity limit both because ofthe fuel specificity ofthe kinetics of auto-ignition and because of the complexity ofthe network of temperature- and pressure-cependent reactions ln the present work, kinetic simulations have been performed using the open source code CANTERA [63,64] and Chemkin-Pro [65}, including the effect of pressure and temperature gradients ‘due to shock attenuation based on the assumption of an isentropic ‘compression behind the reflected shock and a constant pressure sradient of 2.5%/ms. Itisevident from Fig, 4, that even under the “worst-case” uncer tainties deseribed here, significant deviations remain between the simulations nd the experimental results; thats, the potential influ- ‘ence of shock attenuation and pre-ignition on the experimental re- sults is substantially smaller than the deviation between the simulation and the experimental data. Iti also important to note that the deviation between adopted kinetic simulations and the ‘experiments is smaller as the pressures are decreased (10-20 bar) ‘a least for the high-temperature regime as reported elsewhere [47]. This itself suggests thatthe deviation of the experimental data at lower temperatures is caused not by facility effects but by the role ry ie ignition delay [ms] 10: Gor com ies en te 1000/T [17K] Fig. 4 Comparison of kinetic simulation with experimental results considering maximum posible uncertainties due to Taity effect and pregnition elect. Dated tne ints mol vaiation time in Re [44 The ise ares the ‘worst ese stsurpin or the pressure reported 10 0 z & OS rae O oor Omen ean ae 100007 [17K] Fig 5. Comparison of the erature reported models f0 the new experimental results at 0b ‘of reactions that increase at higher pressures and lower tempera tures and that are not included in the previous kinetic models [3438.44.45] A schematic view of the auto-ignition chemistry especially for the reactions dominating in diferent temperature andl pressure re- sgimes is presented in Fig. 6, which is based on Ref. [66]. The scheme reveals that as pressures are increased, the regime where peroxy chemistry as well as the HO, and H;0, chemistry domi- nates (shown in shaded area) extends to higher temperatures. ‘Therefore at higher pressures, the peroxy radical stabilisation is increasingly favoured over redissociation, and its isomerisation ‘could become dominant in providing the crucial branching steps to auto-ignition, The general reaction scheme of the low tempera- ture oxidation is discussed elsewhere (66,67) and only the butanol- specific oxidation reactions in this regime are discussed in detail in the kinetic modelling sections. Small radical chemistry +0, 1200 Hen Tempenue z $ ow 2 HO, and H,0, 7 Chemistry i Inertia Peroxy Chemisty 00, oF 5 0 10 Pressure [bar] ig. 6 Simple general ato ignition epee band 09 Ref. (65 S Vrant eta Combustion and Fame 158 (2011) 1644-1455 449 Additionally, shock-tube experiments at high pressures have advantages over low-pressure ignition experiments because much Shorter ignition delay times are obtained at high pressutes, alow- ing, measurements to be extended to lower temperatures while ‘maintaining measurement times well within the shock-tube re- sgime of validity. Furthermore, forthe same temperatures the inf ence of pressure gradients is significantly reduced with shorter ignition delay times because the absolute change of pressure and temperature is smaller. From the kinetics point of view, the role of peroxy chemistry could also become dominant at elevated pres- sures even at higher temperatures under shock-tube conditions. 3. Detailed kinetic modelling, Several detailed kinetic models have recently been published {34,38,44,45,68,69] which were validated for diverse experimental investigations that have been briefly discussed earlier in this paper. ‘Most relevant to the scope of the current work are the experimen- tal measurements of auto-ignition delay times in shock tubes be- hind reflected shock waves at temperatures and_ pressures Fanging from 1100 to 1800 K and 1 to 8 bar [44,45] and laminar flame speed determinations, extinction strain rates and species concentration profiles in jet-stired reactors [34.38]. These data have been used to validate the reaction kinetic model of n-butanol combustion. Figure 5 shows a. comparison of available literature ‘models |34,38,44,45,68,69] with our current shock-tube exper ‘mental results at the highest pressures. The literature detailed ‘mechanisms overpredict the experimentally determined ignition delays in the entire temperature regime studied here, and the dev ations are even more prominent in the transition from the high- Temperature regime to the intermediate-temperature region. The experiments used to validate these models are at low pressures; ‘one could speculate that at intermediate temperatures and higher pressures the role of peroxy reactions may increase. This shows it is unsafe to extrapolate model predictions to conditions outside of the regime for which the model has been validated. The reaction pathways for peroxy formation from the primary hydroxybutyl radicals and subsequent intramolecular isomerisation are incom- plete or even absent in most ofthe reported detailed models cu rently available for n-butanol ‘The exact cause for the discrepancies between the mechanism and the experiments atthe high-temperature end is unclear, espe- cially considering the differences in the represented chemistry and the “rate estimation rules” adopted by the different groups. One ‘would expect that the high-temperature experimental results, especially at high pressures, would be relatively easier to model since the chemistry dominating at high temperatures is mostly Fuel independent and controlled by small radicals and H + 03 reactions. Moreover, it is a traditional judgment that most reactions would have their high-pressure limit reached at in-cylinder pressures and adopting the high-pressure limiting rate constants is common in detailed kinetic modelling. However, itis true that certain reac- tions, eg. in the H,/Op reaction system have the high-pressure limit well beyond 2000 bar (eg. H+ Oz (+M) > HO3 (+M)) [25] and in most eases the high-pressure limit is not experimentally determined, We do not seek to discuss all individual reactions adopted by the several kinetic models reported in the literature, therefore, ‘we performed a detailed sensitivity analysis especially forthe high temperature regime on the kinetic madel from Black et al [44]. We adopt this mechanism forthe high temperatures because ofits re- atively closer agreement to our results at higher temperatures and lower pressures. The Black etal. [44] detailed mechanism had al~ ready been validated against their own shock tube data (1. up to Sbar and 1100-1800 K) with whieh it agrees quite well, although the authors report non-linear behaviour of the ignition results at lfferent pressures. Comparison with our new experimental data, however, indicates an over-prediction of the ignition times at higher pressures as shown in Fig. 5. ‘The slight pressure and temperature gradients in the shock tube data are accounted for in the simulations by assuming an isentro- pic compression behind the reflected shock. These effects are small, and are represented as uncertainties in the obtained ignition results by employing the “worst-case” assumptions (see Figs. 3 and 4), 3.1, High- and intermediave-temperature kinetic model Relative sensitivity coefficients ofthe mechanism of Black etal {44 are shown in Fig. 7; each rate coefficient is increased by a fac- {or of two and the corresponding change in predicted ignition de- lay time is calculated, The 10 most sensitive reactions are listed for conditions of 20 bar, equivalence ratio 1 and temperatures from 1050 to 1450K. ‘At 1050 K and 20 bar, the reactions to which the ignition delay is most sensitive are the branching reactions of H-abstraction from n-butanol by HO; yielding H.0,, with its subsequent decomposi tion forming 2 OH, and the self reaction of HO» to form HO» and At higher temperatures however, the most sensitive reactions are the reactions of H with 0: to produce OH and 0 and the ther- mal decomposition reactions of the fuel itself. ‘Accurate rate coefficients for certain sensitive reactions (eg. Hal 0. reaction system at high temperatures) can improve the agree- ‘ment for the high-temperature experimental ignition time deter- ‘minations. The most sensitive pressure-dependent reaction from about 1000 K to 1200 K was found to be the decomposition reac- tion of H.0s. 1,0; +M — OH +0H4+M 2 ‘This reaction has been experimentally and theoretically studied by several research groups [70-73]. The most recent experimental ‘measurements involve the combination of shock tubes and UV absorption spectroscopy [71,73]. The rate coefficients used in the model have been updated based on the recent studies and are de scribed by a Troe fall-off expression with the fall-off parameters adopted from Refs. [70-72]. This new rate coefficient expression increases the HO, decomposition rate by 40% at 80 bar, compared 3020100010203 Sensitivity Fig. 7. Sensi analysis on inition delay for tana nition dey a 20 bse for tice diferent tempestues using the adel fom Re. [38 1350 |S ance a Combustion and lame 158 (2011) 1404-1455 to the rate coefficient given in the model of Black et al [44]. At temperatures above 1400 K, the n-butanol decomposition becomes ‘dominant and the axidation reactions of small radical species pro- ‘duced by the n-butanol dissociation play an important role. The ignition of n-butanol at higher temperatures in this regard be- ‘comes fuelindependent especially because of the fragmentation to smaller radicals. The only fuel-specfic reactions inthis T-region are the decomposition ofthe fuel; small radical chemistry, and the reactions in the H2)02 reaction system dominate as Seen in Fig. 7 In the intermediate-temperature region, from about 900 0 1400 K, abstraction of H-atoms from n-butanol by HO» to form 1,02 is a key reaction in the n-butanol auto-ignition. A chai branching pathway is provided by the subsequent decomposition ‘of HQ» to form 2 OH. The n-butanol + HO; reaction to produce H,0; becomes less significant at lower temperatures because of the well defined barrier height and the required activation energy for this abstraction channel. At very high temperatures, it becomes iezelevane as well because of the dominant thermal dissociation route of the n-butanol itself. The rate coefficients adopted in cur- rent models forthe n-butanol + HO, reaction are so far based upon ‘estimates from analogy with alkanes and current kinetic models liffer by a factor of 20 for our experimental conditions [44,43]. This, ‘motivated our new theoretical study of the H-abstraction from n- butanol by H0>, discussed in the next section, which also confirms the major channel leading to a-hydroxybutyl radical and H.02 with decreasing order for the specific rate coefficients to each hhydroxybutyl radical as «>> 8> > OH group. ‘The model predictions, with only the refinements at higher temperatures as discussed above, are depicted in Fig. 8 at 80 bar The agreement with the experiments at higher T suggests that the pressure dependence and accurate fall-off parameters of some 2/04 reactions in addition to the n-butanol + HO; reaction could indeed be very crucial to the intermediate- and high-temperature kinetics. However, itis also true that adopting only the updated high tem- perature rate coefficients for the Ha/Op system and n-butanol + HO Feactions does not improve the agreement with experiments below 1000 K which indicates a different mechanism operating in this regime. This mechanism, also termed the low-temperature ‘oxidation mechanism, is developed and validated for the low= temperature (below 1000 K) experiments at very high pressures ‘and is discussed in detail in Section 3.3, 32. Butanol + HO, reaction ‘The H-atom abstraction from n-butanol by HO> radicals has been investigated theoretically in recent studies which have re- vealed its complexity due to many possible reaction paths [74,75]. Until now the rate coefficients leading to the different product channels in the reaction of HO. with n-butanol had been neither experimentally measured nor theoretically calculated. Modelled rate coefficients for HO, reaction with n-butanol to yield 1,0: and one of the butanol-radicals are based on comparison with H-abstraction from ethanol and alkanes [44], Kinetics vary for the 14 conformers of n-butanol [74,75], as well 2s for the other three isomers of butanol [45]. For each n-butanol ‘conformation there are 10 hydrogen atom abstraction sites with different barrier heights and accordingly different reaction rates, as listed in Table 2 [75]. The conformational study of n-butanol ‘of Ohno etal. [76], using vibrational spectroscopy combined with, ‘ab initio ealeulations, indicated the TGt conformer [76] as the low- est energy form, This conformer has recently been confirmed as the ‘most stable structure by high-level of theory calculations by Moc et al. [77], The notation adopted is that of Ohno et al. [76}, the first Trefers t0a trans conformation about the CC-CC bond, the G refers to a gauche conformation of the CC-CO bond and the final 10 10 ignition delay fms} 10" eee ce eons nanos 100007 [1] Fig. 8 lmition delays at highest pressures, around 80 bar, fr stoichiometric butanol mistres compared 0 deta inte modest Seton 4). able? Barrie heights for H-atom abstraction rom r-btael by Hoy, values assembled from 175) Bs ele wine ° m3 vio es om 3 72 653 3 #3 on 913 lowercase t indicates a trans conformation round the CC-OH bond. Figure 9 shows the numbering of the H-atoms for this lowest-energy conformer, used as the reactant structure in this kinetic study, For our kinetic calculations we used the potential-energy sur face data from Black and Simmie (75]. They report for each hyaro- ‘gen abstraction site G3-calculated values for the barvier heights relative to the lowest 2-7 TS (see fig. 9). The absolute barrier heights are obtained by use of the given CCSD(T)/cc-pVTZ bartier height for the 2-8 TS of 55.3 kJ/mol including MP2 zero-point en- ‘ergy (ZPE) with 6-311G(d.p) basis set [75], The resulting values {are given in Table 2. A common approach to reduce complexity is to treat hydrogen atoms that belong to the same carbon atom equally. In this study however, we calculated all 10 transition as . OH 3 8 a os 9 Fig. 8. The most stale n-butanol conformer, Tt rd numbering of he atoms, the bon eng an angle are vena Sepptingilormaton ‘Veal et Combustion and ame 1582001) 1444-1455 451 sates separately and then summed up, per carbon atom, the rate Coefficients when presenting the results for H-abstraction by HO, {tom each C-atom and the OH in n-butanol. In order to apply transition-state theory, one has to calculate the partition functions and therefore determine the vibrational fre- quencies. We calculated anharmonic frequencies (as well as har- monic ones) using Gaussian03 [78]. Because of the good performance of B2LYP with tzvp basis set concerning geometry and frequency calculations {79], we re-optimised the MP2 geome- tries from Black and Simmie [75] and computed the frequencies from these new geometries. These frequencies are given as Su porting information. The torsional modes have been treated like vibrational modes in conjunction with anharmonic fundamental frequencies. For abstraction from a carbon atom, the imaginary fre- ‘quencies are quite high (~1600i cm '), resulting from a large cur- vature atthe saddle point, Thus the shape of the free energy curve should be dominated by the potential energy and variational ef- fects should be small. Tunnelling might have more influence in- stead, because the large curvature causes the barrier to be We used the G3 barrier energies, including ZPE obtained from frequencies computed with the Hartree-Fock method, from Ref. 175] ta base our kinetics on the same adiabatic potential energies as in this reference, We used the POLYRATE 2008 codes [80] to ‘complement simple transition-state theory with tunnelling factor computed along a straight-lined reaction path (ZCT-meth zero curvature tunnelling (81), described by a fitted Eckart poten- tial, The expression forthe rate coefficient reads eros HO) Fe das 2 (- aT 6 (Cis =Gu—Ge : Kp exp (8554-2) 3) ‘With the standard concentration c= 1m? mol * and. c Hs 0 (i) (Gri) ) ‘The influence of the width of the Gibbs-free-energy barrier on the rate thus becomes more dominant as the temperature in- creases, as the effect of energy differences decays with increasing, {temperature while the effect of entropy is sustained. We can thus evaluate neither rate coefficients of the different channels nor of the different hydrogen atoms of one group based solely on energy information at 0 K. An accurate modelling of branching ratios must include energy and entropy information ofall the transition states Entropy effects are thus accurately taken into account for each H- atom separately Furthermore, entropy effects depend mostly on the low fre~ (quencies, which are strongly dependent on a detailed treatment of anharmonicity. An accurate quantum-mechanical determination ofall frequencies, including the imaginary frequencies is a crucial task in this study. The partition functions are thus computed based on these anharmonic frequencies. In the Gaussian03 package {78}, with which the anharmonic frequency calculations were per- formed, the anharmonic frequencies ae included in the harmoni ‘oscillator partition functions. This has been proven to give good results for TST calculations (82 ‘The calculated rate coefficients for H-abstraction at each atom and from the OH group by HO2 are presented in Fig. 10 The specific rate coefficient expression leading to the different rad- icals are given as follows (units cm? mols ") ke = 1.58 « 10°" TP exp(-5289/7) ky = 6.48 x 10°? T exp(-7550/P) 10" rr w 10 1 wo .. Sy ww . a0! 07 08 08 40 41 42 43 44 0007" [14 Fig 10. The calculated HO,* n-butanol rate coefficients for the diferent radical formation channels. The channel 02a early dominates ever thi temper ture range fom 7009 1400 k, = 5.89 10? T* exp(-6618/T) ky = 1.25 « 10* 71 exp(-8548/T) oy = 3.78 x 10" > exp(-9617/T) ‘The dominant channel in the intermediate-temperature region {for n-butanol ignition, roughly from 900 up 10 1250 K, is clearly the formation of the a-radical: more than 60% of all hydroxybutyl radicals formed in the HO, + butanol reaction are a-radicals. For- ‘mation rates off y- and f-radical are about 5-10 times slower than a-formation in this T-tegime, while H-abstraction from the hydroxyl group is 100 times slower, as this H-atom is the most strongly bound and this channel was expected to be negligible. The importance of the HO; + n-butanol reaction strongly declines at relatively high temperatures (above 1400 K) due to its competi- tion with the thermal decomposition of n-butanol itself. 3.3, Low-temperature model generation Auto-ignition in the low-temperature regime is critically sensi tive tothe initiation oxidation chemistry that occurs via the forma- tion of peroxy radicals (RO;) by reactions of primary fuel radicals ‘with > [83-85]. This peroxy chemistry has recently been intensely studied for several alkanes [26,86-88] and the oxidation of ethanol {67.89-91], The RO, radicals can either redissociate, or eliminate HO, or OH, or isomerise to hydroperoxyalkyl radicals (QOH) The QOOH species itself can also decompose to bimolecular prod- ucts including OH and HO>, More importantly, the addition of an extra O; molecule to the QOOH is proposed to be the major chain-branching. pathway at the low temperatures: 02000H ‘somerises to give a radical with two hydroperoxy Functional soups that can decompose to eventually yield two OH radicals. These pathways from hydroperoxyalky| radicals have been pro- posed for a long time and new evidence for such a mechanism hhas been reported by detecting the ketohydroperoxide [92] which {san important intermediate species in the chain-branching path, even though the QOOK still remains undetected. Higher pressures stabilise the RO» and QOOH species, and higher temperatures lead to destabilisation making this a very complex network of reactions as a function of temperature and pressure. As temperatures are in- ‘creased, the thermal dissociation of the QOOH becomes faster us. |S ance a Combustion and lame 158 (2011) 1404-1455 Which is one reason for the NTC region. A general scheme of the reactions participating in the low-temperature oxidation is pre- sented in several other publications (66,67 92] and will not be dis- ‘cussed in detail here. Other OH-yielding reactions that may contribute to the ignition are the reactions of HO, with R yielding two OH radicals and the reaction of RO2+HO: to form (OW + products Ignition in the low-temperature vegion thus critically depends, ‘on the strongly temperature and pressure-dependent rate con- stants forthe reactions involved in this peroxy chemistry, leading to branching through the proposed QOOH +02 channel and Re+HOp channel, termination by HOp elimination from RO, and (QOOH and decomposition of R, or chain propagation by dissoc tion yielding OH. A full n-butanol kinetic mechanism for low tem- perature would require knowledge of the details of the peroxy radical chemistry for butanol. Tis is currently the subject of com- putational and experimental investigations in our (and other) lab- ‘oratories, but for the purposes of demonstrating and exploring the role of the low-temperature chemistry inthis system we derive an ‘ad hoc “simplified” scheme of the peroxy sub-mechanism by ‘employing analogies to ethanol oxidation, which has already been studied in detail [67). This simplified scheme is summarised in Ta- ble 3 and has been added to the refined high-temperature mecha- nism that has been discussed above, ‘We have used the term “simplified” because this mechanism is based on ethanol oxidation, which involves only two primary alkyl radicals, the 2- and f-hydroxyethyl radicals, whereas in the case of butanol there are four distinct primary radicals. The rate coeff- cients for the ethanol-based peroxy reactions [91] successfully modelled OH-LIF measurements of pulsed photolytic Cl-initiated ‘ethanol oxidation that will be published elsewhere [93]. The sim- plified model is clearly an interim measure, giving a “Iumped” per- ‘oxy chemistry that is consistent with analogous systems. Because the low-temperature oxidation kinetics of n-butanol is not yet well investigated at a very fundamental level, itis still necessary to investigate the initiation elementary reactions of n-butanol oxida- tion to obtain accurate rate data. The biggest challenge may lie in understanding of the subsequent chemistry leading to QOOH spe- ies and low-temperature branching due to the second 0» addition to the QOOH, The second Os addition and its rate data still remain uncertain, Theoretical investigations on this crucial reaction and the characteristics of its potential-energy surface (PES) have just begun and predictions fora few alkanes (propyl, 2-pentyl, neopen- tyl, cyclohexyl) have been reported in the literature [2686-88], In the present sub-mechanism, we adopt the ad hoc rate coeffi- HO» +CxHFOH, an unsaturated alcohol, is lowered by a factor of 5 a indicated in Table 3. This modified rate is much ‘loser to the cyclohexane reaction system. For a more accurate ‘deseription, it will be important to probe the PES for the initial oxi- dation steps of n-butanol oxidation supported by detailed theory, ‘experiments and modelling, Such high-level investigations are no doubt forthcoming, but introduction of this analogous ethanol- based mechanism into detailed kinetic models may prove helpful in the interim, 34, Comparison to experiments and sensitivity analysis Figure 8 compares the experimental results to the new model, predictions which now includes the complete mechanism for both high and low temperatures and compared for pressures up to 80 bar. The new detailed model gives a good agreement with the ‘experimental results which also includes the maximum uncer tainty caused due to facility and pre-ignition effects. This work provides further motivation for more experiments at temperatures Compl istof reactions ince in the simplified mode! for low-temperature proxy chemistry. The notation *QOOH" refers oa lumping ofl hyéroproxybuty radial Isomers. Rate coefclets ae based on analogy with perory chemistry rections of ethanol 313] and alkane axdaton schemes as lacssed i the text. These peau ‘dependent rite coins ae sve fora pes of 0 b3 Uns: moles cal ma Alp) and -hydronybiy radical (CaOH) re tented enti fo thi penny sub-mechanism. C/OH is an unsaturated acehel or butanol depending on he primary radical More deta reve in Supporting information. GiHyOH =O, OHO, Tote ae 255 (9195) EHOH-0, » CPN 0, 201E33 sm 3550 {o193}, yoti-0, = GHeCHO + Cro oH, Sasa eae 35960, {91951 Eo 0. 00H ome 727 mm {o193}, CHHOH + Ho, 20H + caoH oaTe-12 ° ° {193}, GHIOH-0, gooH ‘owe aa soz {or93} GHYOH-0; = HO; = caon eos 25 mas {193} CWOH-O, + HO, + OF © 0, product, Taseon1 ° ° {193}, (Qo0H +0, > OH volt product, Soxesor2 ° ° [esi QooH On FcHLOCaRseHO Tonks aan Tso {o193) GOH 0, HOH + HO, inet an saa (193) ‘Veal et Combustion and ame 1582001) 1444-1455 us ‘much lower than those investigated in the current work as well calls for accurate theoretical models for the low-temperature oxi- dation kinetics especialy for elevated pressures. Sensitivity analysis in Fig. 11, reveals that at high pressures and low temperatures, 80 bar and 800 K, the most sensitive reactions are the reactions forming the peroxy (RO) from K+ Os, RO to form HO, + products, RO» isomerisation to QOOH and the back dissocia- tion ofthe RO, to form R + 0,. The second 0, addition to QOOH is a critical step to branching and is indeed very important but since the assumed rate constant for this reaction sequence is high, it does not appear in this analysis. Rather, the rate-cletermining step for branching is mostly the isomerisation of RO, under these assumptions. As expected, the n-butanol initiation reactions by ff and £10», yielding the primary radicals that subsequently form the peroxy species, prove to be sensitive as well In the same figure, the results ofthe sensitivity analysis of the ‘new model for 1, 10 and 40 bar at 800 K are also plotted, One dis- tinctly observes that the reactions sensitive at 80 bar are far less sensitive at lower pressures especially at 1 bar. This reflects the fact that high pressures greatly enhance the stabilisation into the RO, and QOOH wells. This further suggests that chain branching may be favoured at high pressures because of stabilisation into the QOOH wells. The isomerisation between the RO» and QOOH species may be much faster than QOOH dissociation and this could enhance chain- branching in low-temperature conditions. Under “teal engine” conditions close to 100 bar the reaction, sensitivity and the relevant chemistry is quite different from that at low pressures. Therefore, pressure dependence isa very signifi ‘ant factor. This is a challenge to kinetic modelling because both product branching ratios and the role of peroxy reactions can be strongly pressure-dependent and reactions that are found insignif- ‘cant at low pressures may be very influential in determining ig tion characteristics of novel fuels at high pressures enatsrajecnat 0, cnoea.=cnanasro, mar soe Shawegacnoette, Ee so Nancdocieeton Berner epincy oa Seer renrcnal= enone menaonee tenonsne, cae, rene ngrmeravee ‘05 04 03 02 04 Sensttty 00 07 02 03 04 Fi. 11 Sensivity analysis withthe presented »-butanol detailed mechanism ‘Mlostrating the influence of pressure on inion dey and dominant reaction pathways The sensitviy factor reported Is the fraction change in aulated fenton delay pon doubling the ete coticet of angle recon. Our new model with the refined high-temperature chemistry nd the simplified scheme of low-temperature peroxy reactions ‘was further compared with our previously published shock tube ignition delays at 20 and 40 bar [47] as shown in Fig. 12. We also compared our model predictions to the available literature re- ported data of Moss et al. [45] for the low pressure data in Fig. 12. For all temperatures and pressures, our detailed kinetic ‘model is in good agreement with the experiments 3.5, Predicted species profes Figure 13 depicts the species concentration profile as a function, ‘of time for crucial species identified through the sensitivity analy- sis given in Fig. 13 for 800 K at both 80 bar and 1 bar. The ignition is identified after 22 ms, coinciding with the consumption of the fuel and prompt production of a large amount of OH radicals. The radicals (R) produced by H-abstraction either add to 0» to form RO» oF decompose, The concentration of both R and RO, builds up until the onset of the peroxy chemistry after 0.5 ms, followed by an increase in fuel consumption. The QOOH concentration is still an order of magnitude lower than the radicals (R) concentration and its mole fraction peaks at a value of only 3 » 10 *; its fast subse- ‘quent reactions with O, prevent its concentration from building up. Its short kinetic lifetime and low peak concentration also ‘makes QOOH difficult to detect in experimental studies. As ob- served in the species profile and discussed in detail above, the cru- ial step to branching remains the RO, isomerisation to QOOH. The ‘ther species in higher concentration is the HO, which is much higher than the OH concentration prior to ignition. This eventually changes as the HO» is very rapidly consumed forming OH causing ‘ignition, Figure 13 also gives the species profile of the new model at 800 K and 1 bar. The sensitivity analysis in Fig. 1, for these con- itions indicates that the peroxy chemistry Is not prominent in influencing the ignition under these conditions. This is confirmed by this species profile, showing a long ignition delay of more than 30 ms and negligible production of the peroxy species, 4. Discussion Both experiments and kinetic model reveal the emergence of significant role of the peroxy chemistry at very high pressures ‘which is far less prominent even at modest pressures of 10 bar, Although minor deviations stil exist between model predictions Dba ge @, we xh sD bar 06 07 08 09 10 14 12 13 1000771126 Fig. 12. Comparison of inition delays at diferent pressures with Kinetic simul wast |S ance a Combustion and lame 158 (2011) 1404-1455 ous ool SAE 200% bar \ 2 o0s \ q \ 7 om : ont time {ms} time [ms] Fig. 13, Species time profes for ignition on butanol at 800K and 80 bar clelated with ee presented detaed mechanism. and experimental results, especially at the lower end of the inves- tigated temperature range, suggestions of the NTC behaviour are “observed as pressures are increased, Around 900 K the ignition de- lays remain almost constant with decreasing temperature and start to increase again as temperatures ate lowered below 830 K. The ‘curvature observed in experimental results is well captured by the model, although this curvature is slightly different in the pre- dictions below 900 K, ‘The deviations between the model and the experiments could arise from the simplification in the reaction scheme adopted here, Which is primarily based on ethanol oxidation, The discrepancies between the madel and the experiments could probably be Further reduced by a more detailed inclusion of the low-temperature ‘chemistry based on accurate theoretical andjor experimental tud- ies of R with O, RO: and QOOH reactions for the n-butanol. This modelling study suggests a high priority for further elucidation of these crucial reactions. Furthermore, the reaction of QOOH with, (02 to yield two OH radical isa critical reaction sequence for which ‘accurate information is lacking. In this model 2, f-. y= and &- hhydroxybuty! radicals formed by H-abstraction from n-butanol ‘are considered the same for their peroxy chemistry. Although this work guides further theoretical research, the ‘experiments are also performed in a very limited temperature re- sgime. In order to validate the peroxy sub-mechanism and to an- swer some of the open questions regarding the uncertainties in ‘experiments and also on the accuracy ofthe sub-mechanism kinet- ics, more experiments would be needed below 795 K. New exper- iments are currently underway employing altemative facilities ‘which will extend this regime to focus more on the low tempera ture combustion aspects especially at very high pressures for ‘butanol oxidation. 5. Summary and conclusions Ignition delays were measured in high-pressure shock tubes for the temperature regime of 795-1200K at pressures from 61 to ‘92 bar. The aim of this work was to understand the kinetics of igni- tion for low-temperature combustion of n-butanol for its applica- tions to advanced engines. Both experiments as well the newly developed kinetic model strongly favour the influence of peroxy ‘chemistry in the ignition of n-butanol at elevated pressures. Increased pressures indicate increased influence of NTC causing, hon-linear “up-turns” as the oxidation transits from the high- temperature to the intermediate-temperature regime, which is a ‘ypical signature for NTC behaviour. The detailed kinetic model for n-butanol has been developed based on refining the high tem- perature kinetics from existing reaction mechanism and introduc- ing a new reaction scheme for the low-temperature ignition pertinent to the peroxy chemistry. Complete model shows good ‘agreement for ignition delays with current work and also earlier Published work in the literature. New experiments for stil higher pressures and lower temperatures will help to validate the model This work also guides further research in the theoretical investiga tions for the elementary initiation reactions of peroxy formation ‘and isomerisation for n-butanol oxidation. ‘Acknowledgments ‘The authors thank Cluster of Excellence “Tailor Made Fuels from Biomass”, which is funded by the Excellence Initiative by the ‘German federal and state governments to promote science and re- search at German universities. CA.T acknowledges support from the Division of Chemical Sciences, Geosciences, and Biosciences, the Office of Basic Energy Sciences, the US Department of Energy, as part of the Argonne-Sandia Consortium on High-Pressure ‘Combustion Chemistry. Sandia National Laboratories is a multipro- ‘gram laboratory operated by Sandia Corporation, 2 Lockheed Martin Company, for the United States Department of Energy, under Contract DE-ACO4-94AL85000. ‘The authors thank Dr. Judit Zdor and Dr. Oliver Welz fr help- ful discussions and for sharing certain data prior to publication, Appendix A. Supplementary material ‘Supplementary data associated with this article can be found, in the online version, at doi:10.1016/j.combustflame2010.12.028. References 111K. Kotse-Hoinghaus,P Osswak, TA Cool Kasper 8. Hansen, FQ. CX: ‘Westbvook, Pt Wesimotcand, Agen Chem nt Ed 49 (2010) 3572-3597, (a) 1th uc HF Bak iOpen 1200 20-2 Ha] S.Asumi F Hana [C-Uo, Nae 451 2008) 36-013, [5] YICUin IP. Basco App Enron. Microbiol 45 (1983) 966-973 (6) Ezeh. Qureshi, HP Blache. Biotechnol. Bioeng. 97 (2007) 1460-1458. [7] AK Agarwal, Pog Energy Combs Sek 33 (2007) 233-271 [sf A‘Deribs, og ergy Combust Se 333007) 1-18 ‘Veal et Combustion and ame 1582001) 1444-1455 55 [ol 4. chen, SJ Sal XC Wang Proc Combust est 31 (2007) 2961-2589. [HO] Re Lapaet, ema, Rade Fenanden Trg. Energy Cambust Sk 24 (2008) 198-223, [7] Seatac twamoen, savas, M Yamada. Kanko, Poe. Combust. Inst 51 ow) sist 3138 [12] €0. tole, Vi, McBone Proc. Comiust.nst 32 (2008) 2049-2056, [131 TG fang. CFF ee Pro. Combos 32 (2000) 2785-2792. UAL IM. Anderoh. Ac Piper AP. a Croz R Bounaceur F. Batn-ecec Dagaut X Montagne, Pro Combust- mst. 52 (2000) 2851-2850 (15) EJF Hata Douwel Aj: Donkerarok, AP van Vile, NUD. Bost LAL. Somers RSG. Bure NJ Dam, ter Meuien, Proc. Combust inst. 39 (2005) 2817~ 2525, [161 Bina M, Sods, S. Kir, K. Yass. H, Suk Hoda, S, Yamamoto, S (king SuzuklH. Yukawa, Appl MierbioLsorechnl 7 (2008) 1305" [17] &xalschouer.T.Stldng A sttnbucel. Mroblology 152 (2006) 2520-2536. [is] St. withers, SS. oct, Ley, JB. Newman, JD. Keating, Appl Envcon. ‘Mirobil 73 (2007) 6277-0283 [19] 1D easing, ACS Chem. Biol. 32008) 64-76. [2a] Vip. rua, bj, Pera St. Withers 13. Newman, JD. Keating, Nat Bote 31 (2003) 798-802 [21] DK Ro, EN Paradise M Oui, Kiser, KL Newman, JM. Nduoge, KA, Ho, RA Eachus, TS. Ham Kit. MY. Chang. ST. Withers. ¥. Shiba, Spon 0. Reasling, Nature 40 (206) 840-345. 22] wor ebro, 2010 [23] woe fuccenerswth-aacende, 2010. (lpn inmates 0207 ene [25 BX. Femandes, Kher, Tro, WG, Ushakow, Phys: Chem, Chem. Ps. 10 (2008) 4513-4331 (26 RX Femandes J Zidor, LE Jasin, JA. Miles, CA Tastes. Phys, Cher, (ent Py (200) 1320-1527, [20] RX Fernandes, Ker, j Tie, Phys Chem. A 114201) 2962-9968, [28] RV. Nav GN. Benet ET Faposaks J Bacteriol. 176 (1984) 871-885, [25] sentheshanmaganathan, lochem 7 (1960) 568-575 [60] ¥. Tash, H shit, Ne Hoyas, Baba Kobayashi, K Sonoma. ls Beng 168 (2007) 238-240. [B11 YTashir. KC Takeda, G.Kabayasi K Sonomot. J. Biotechnol 120 (2005) yor-208, [52] N-uresh 1s. maddox J Ferment. Boe. 80 (195) 185-189, [35] Lo-Keasing chou St. ltechno 29 (2008) 26-208 [34] PS. Veloo, VL Wang FX Egolfopodios CK. Westbrook Combust. Fame (2010) di10,0.6 combs 20100401, [351 B Yang P. Oswald YL} Wang, UX: We, ZY, Tian, FO K: Kobe Hoinahous, combust. le 148 (2007) 198-206, (36) up fee butane co, 2010 [39] Cs eel... Pete Poe. Combust ast 30 (2005) 1365-1570. [B31 SM, Srathy” MJ. Thomson, ¢ Torbé P.Dapaut, F. Halter, C. Mounaim Roussel, combust. Pame 156 (2000) 852-804 [39] Kasper U. Strckmee, Oss. KKose-Hoinghaus, Pr. Combust Inst 32,2008) 1255-1202. [do] P Osswald, HColdenbors X Kohse-Hbnghaus,B. Yang, T. Yuan, FQ, Combust ane (2010, doe10. 1016) combs 201006002 (411 JA Barard Trans. Faraday Soc $3 (1957) 423-1430. 14a] TS. Norton, L Dye Pot. Combust ns. 23 (1991) 179-185, [23]. Dagaut SM. Sarath, My Thomson Proc. Combust Inst. 32(2008) 220-237. [4] G lack Cran, Pichon. IAT Sn . Zuko, Combust. ame 157 (Goo) 363-375. [a5] [T Ms, AN Berkowitz, MA. Oehlchaeger | Be, V. Wah, PA GlaudeF fatn-ecer Phys Chem, A112 (2008) 1883-10855, [46] $S" Vasu, DE Davdzon, Rc Hanson, DM. Golden Chem. Phys et. 497 (2010) 26. [47] KA Healer, RX Fernandes 1 Otie,J Beckmann 0. Roel. Peters, ae Combat ns (2010), dae0 1016 gro 201006052 (4) FAC Alsou, App Therm Eng. 17 (1997) 537-548, (9) EMC Alsou nt) Energy Kes 21 (187) 21-0. 150] Mtn Man 0, Crs, Pa oe. Mk Eg Pt 424 20) {51] M. Gautam, D. Martin I ro. Inst Mech Eng ar A 24 (200) 497-51 [52] Fic Bhatachry,s.ehaterees FN: Mishra, App En. Nene 20 (2008) 252- [52] ¥ Zhang, AL Roshan, Combs, ame 157 (2010) 1816-1824, [54] K Heute. Olver Shock Waves 20 (2010) 307-116. [55] DE. Davison. RK: Hanson In. cher, Kine. 36 (2004) 510-523. [56] EL retersn. ek Hanson, shock Waves 10 (2001) 05-20. [52] GA tang DF Davidson Ka Paso, Pot combust Int. 12 (2008) 181-188. [Ss] MSP. tahundawala SAP. Coreta 5 Wiss, CD. Ca, SS. Su, Mt ‘hem. Kinet 38 (2006) 194-201, [59] Rumenthal Kc Feweger. KH. Komp. . Admit. Combust, Sc Technol 113 (95) 137-168. (60) BL Wang Over, .cronig, combust Fame 135 (2003) 93-105. [et] HL Ze Owens DE Davidson, RAC Hanson ne J. Chet. Kinet 40 (2008) [62] M. chaos FL Dryer, Inc. Chem. Kinet 42 (2010) 43-150, [63] tp: sourefonenetpietfentera 2010. (a D- Goodwin, Cantera: Object-inted Stare for Reacting Flows, Technical Report, Calo lnsitute of Tecnology, 2005 {65} Chemkin Pro 1509, Reaton Design, a Bo, 2000 [66] JA Mier Pin Troe. Pro: Combs. ts. 30 (2005) 43-88. [691 Zader, CA. Tastes. RX. etnases, Pog. Enemy Combus. Se, (2010, foi: 101016) pees 201006006 [6] Gran, A Frassoldat. aravell U. Nieman, E Rana, R Sis, Catia, IC Seanad, Combast Fase (2010), do 101016) combastlane 01005 0 [6] Mie Hep, Ki Van Geers, SP- P91, Mari, Wa Gren, Combus Fae 2010.in press, dt 10,1016). combustlme 201006002 (70) SR Sate Georges JA Mil. Phys. Chem. A113 (200) 4457- 171 2A og ra ue DE Don RX Han PCa A {72} | Troe. V6. Ushakoy Phys Chem. Chem. Phys. 10 (2008) 3915-3824. [75] © Kappel other}. Toe. Phys Chem: Chem, Phys. 4 (2002) 4392~598 (lM iJ Smif Cran comp Mods Se 175] G Blick. LM. simi. Comput, Chem, 31 (2010) 1236-1248. 16] Cofino Yoshida. H, Watanabe. Fut, HL Matsira, J Phys. Chem 98 (hs 246030 (27) soci Siam, Hy. Curan, Mo. struct. 928 2000) 149-157 {fal A rite, etal Caustands Cussan in, Wallagord Cl 2004 [79] LA Montgomery. My. rth, W,Ochtesi GA Petersson. | Chem Phys. 110 (999) 28229877 [30] | henge al POLYRATE 2008 [st] 4 Kuppermasn 06. Tue. Am. Chem Soc 95 (1971) 1840-1851 [aa] Vtarone.} chem. ys 120 (2008) 3058-3065, [331 CKC Westbrook Proc, Combust Ins. 2 (2000) 1363-1577 [84] G Melon P. Zou, Si, Klppestein, Me Ahmed, SR. Leone, CA. Totes, DL Osborn Am, Chem oe 128 (2008) 13358713507 Us] in a Sly Gay, rane, DI Csba CA Tae An. (6) Li Hay, HH caratensen, AMM Dea, J. Phys, Chem. A114 (2010) 6584- ‘07 (37) AStran, LW. Borzell | Ps. Cher, A 114 (2010) 7633-7708 [33] H.Son JW Bozzi Phys Cem. A 08 (2004) 1694-1711 {sol J 2idor, RX femands, LE hsm JA Miler CA Tass, ree Rade Symposia, Stanians, Flan, 2005. [90] Ri Ferman) Zidor LE juss, JA. lie, CA Tans, US Combustion Meeting. Ann Arbor, 2000 [01] L280 Remandes,¥-eorgcwski Meloni, CA Tate 1A Mills, Pro. (Combuse nst 32 (2008) 271-277 [02] Fat Lecter 0. Reebne, PA Claude, Fours, 2¥- Zhou, LL Deng (Guo Me, Ange Chem, Il 49 (2010) 3169-3172 (93) J Zidr. RX. Femandes LE. Jui }A Miler CA. Tats, n preparation

You might also like