+ All Categories
Home > Documents > Low temperature auto-ignition characteristics of...

Low temperature auto-ignition characteristics of...

Date post: 01-Apr-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
11
Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend fuels: Ignition delay time measurement and kinetic analysis Yang Liu, Chenglong Tang * , Cheng Zhan, Yingtao Wu, Meng Yang, Zuohua Huang * State Key Laboratory of Multiphase Flow in Power Engineering, Xi'an Jiaotong University, Xi'an, 710049, China article info Article history: Received 29 December 2018 Received in revised form 31 March 2019 Accepted 20 April 2019 Available online 24 April 2019 Keywords: Methylcyclohexane Ethanol binary fuel Rapid compression machine Ignition delay time abstract New ignition delay times (IDTs) of binary methylcyclohexane (MCH)/ethanol fuels were provided by using a rapid compression machine. Results show that MCH exhibits typical negative temperature co- efcient (NTC) regime bounded by two turnover temperatures (T upper and T lower ). As the ethanol blending ratio increases, IDTs increase, NTC regime shrinks and shifts to the lower temperature. Measured data are then used to validate the most updated kinetic model (Bissoonauth et al. Proc Combust Inst 2018), which well captures the IDT and NTC behavior dependence on the ethanol blending ratio. The kinetic reason for the experimentally observed ignition behavior is that ethanol addition reduces the low temperature heat release and slows down the temperature rise and H 2 O 2 decomposition pathway activation. The mole fraction of ROO and QOOH decrease rapidly due to the reduction of MCH initial concentration, and the NTC behavior becomes less apparent consequently. Further comparison among the ignition behavior of the different binary fuels have been conducted to access the effect of ethanol addition on different structure fuels. Results indicate a generalized effect of ethanol addition and the kinetic reason is that the presence of ethanol just simply reduces the hydrocarbon concentration without fuel-to-fuel interactions. © 2019 Published by Elsevier Ltd. 1. Introduction Biofuels are of increasing interests as alternatives for petroleum- based transportation fuel substitution because they offer the long- term promise of carbon neutral and reduce climatic impact. Ethanol is one of the most common biofuels and has been widely used in gasoline and diesel engines [1 ,2]. Because ethanol contains 35% oxygen, adding ethanol to transportation fuel results in more complete combustion and investigations have showed that ethanol addition reduces the soot precursor emission [3e5]. Besides, ethanol is also used in homogeneous charge compression ignition (HCCI) and reactivity controlled compression ignition (RCCI) com- bustion engines because of its high octane number, which makes the regulation of the strong pressure gradient and optimization of the two-stage heat release possible [6e8]. The addition of ethanol to fuel affects certain key properties that affect their use in engines such as the blend stability, viscosity, and lubricity. In addition, the ammability of blended fuel and the unregulated emission prod- ucts such as aldehydes limit the ethanol content of blend. As a consequence, the effect of ethanol addition on the fundamental combustion parameters and the oxidation kinetics of the fuel blend need to be thoroughly understood for its application in practical engines. An accurate chemical kinetic model can be used to predict important engine combustion parameters such as pressure rise rate, heat release rate and intermediate species proles and these parameters are helpful, in terms of chamber design, fuel injection and ow eld organization. It is also important for the develop- ment of clean-burning and energy-efcient combustion tech- niques. As a consequence, extensive investigations have been conducted to understand the effect of ethanol addition on the global combustion parameters and oxidation kinetics of gasoline, diesel or kerosene surrogates, such as n-heptane and iso-octane, which respectively represents the main straight chain and branched chain surrogates. Specically, Lipzig and co-workers [9] measured the laminar ame speeds of n-heptane, iso-octane, ethanol and their binary and tertiary mixtures. They showed that the laminar ame speeds of the n-heptane/iso-octane mixture (50/ 50, Vol. %) equals to the average of the laminar ame speeds of n- * Corresponding authors. State Key Laboratory of Multiphase Flow in Power Engineering, Xi'an Jiaotong University, Xi'an, 710049, China. E-mail address: [email protected] (C. Tang). Contents lists available at ScienceDirect Energy journal homepage: www.elsevier.com/locate/energy https://doi.org/10.1016/j.energy.2019.04.132 0360-5442/© 2019 Published by Elsevier Ltd. Energy 177 (2019) 465e475
Transcript
Page 1: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

lable at ScienceDirect

Energy 177 (2019) 465e475

Contents lists avai

Energy

journal homepage: www.elsevier .com/locate/energy

Low temperature auto-ignition characteristics of methylcyclohexane/ethanol blend fuels: Ignition delay time measurement and kineticanalysis

Yang Liu, Chenglong Tang*, Cheng Zhan, Yingtao Wu, Meng Yang, Zuohua Huang*

State Key Laboratory of Multiphase Flow in Power Engineering, Xi'an Jiaotong University, Xi'an, 710049, China

a r t i c l e i n f o

Article history:Received 29 December 2018Received in revised form31 March 2019Accepted 20 April 2019Available online 24 April 2019

Keywords:MethylcyclohexaneEthanol binary fuelRapid compression machineIgnition delay time

* Corresponding authors. State Key Laboratory ofEngineering, Xi'an Jiaotong University, Xi'an, 710049,

E-mail address: [email protected] (

https://doi.org/10.1016/j.energy.2019.04.1320360-5442/© 2019 Published by Elsevier Ltd.

a b s t r a c t

New ignition delay times (IDTs) of binary methylcyclohexane (MCH)/ethanol fuels were provided byusing a rapid compression machine. Results show that MCH exhibits typical negative temperature co-efficient (NTC) regime bounded by two turnover temperatures (Tupper and Tlower). As the ethanol blendingratio increases, IDTs increase, NTC regime shrinks and shifts to the lower temperature. Measured data arethen used to validate the most updated kinetic model (Bissoonauth et al. Proc Combust Inst 2018), whichwell captures the IDT and NTC behavior dependence on the ethanol blending ratio. The kinetic reason forthe experimentally observed ignition behavior is that ethanol addition reduces the low temperature heatrelease and slows down the temperature rise and H2O2 decomposition pathway activation. The molefraction of ROO and QOOH decrease rapidly due to the reduction of MCH initial concentration, and theNTC behavior becomes less apparent consequently. Further comparison among the ignition behavior ofthe different binary fuels have been conducted to access the effect of ethanol addition on differentstructure fuels. Results indicate a generalized effect of ethanol addition and the kinetic reason is that thepresence of ethanol just simply reduces the hydrocarbon concentration without fuel-to-fuel interactions.

© 2019 Published by Elsevier Ltd.

1. Introduction

Biofuels are of increasing interests as alternatives for petroleum-based transportation fuel substitution because they offer the long-term promise of carbon neutral and reduce climatic impact. Ethanolis one of the most common biofuels and has been widely used ingasoline and diesel engines [1,2]. Because ethanol contains 35%oxygen, adding ethanol to transportation fuel results in morecomplete combustion and investigations have showed that ethanoladdition reduces the soot precursor emission [3e5]. Besides,ethanol is also used in homogeneous charge compression ignition(HCCI) and reactivity controlled compression ignition (RCCI) com-bustion engines because of its high octane number, which makesthe regulation of the strong pressure gradient and optimization ofthe two-stage heat release possible [6e8]. The addition of ethanolto fuel affects certain key properties that affect their use in enginessuch as the blend stability, viscosity, and lubricity. In addition, the

Multiphase Flow in PowerChina.C. Tang).

flammability of blended fuel and the unregulated emission prod-ucts such as aldehydes limit the ethanol content of blend. As aconsequence, the effect of ethanol addition on the fundamentalcombustion parameters and the oxidation kinetics of the fuel blendneed to be thoroughly understood for its application in practicalengines. An accurate chemical kinetic model can be used to predictimportant engine combustion parameters such as pressure riserate, heat release rate and intermediate species profiles and theseparameters are helpful, in terms of chamber design, fuel injectionand flow field organization. It is also important for the develop-ment of clean-burning and energy-efficient combustion tech-niques. As a consequence, extensive investigations have beenconducted to understand the effect of ethanol addition on theglobal combustion parameters and oxidation kinetics of gasoline,diesel or kerosene surrogates, such as n-heptane and iso-octane,which respectively represents the main straight chain andbranched chain surrogates. Specifically, Lipzig and co-workers [9]measured the laminar flame speeds of n-heptane, iso-octane,ethanol and their binary and tertiary mixtures. They showed thatthe laminar flame speeds of the n-heptane/iso-octane mixture (50/50, Vol. %) equals to the average of the laminar flame speeds of n-

Page 2: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

Fig. 2. Experimental repeatability verification and definition of measured IDT.

Table 1Compositions of the tested mixtures.

Fuel f Tc/K Pc/bar

E0 0.5, 1.0 690e920 20E25 0.5, 1.0 680e920 20E50 0.5, 1.0 680e930 20, (10, 30 for f¼ 1.0)E60 0.5 700e920 20E75 1.0 680e860 20E100 0.5, 1.0 850e890 20

Y. Liu et al. / Energy 177 (2019) 465e475466

heptane and iso-octane at the same condition. However, blending50% ethanol into either n-heptane or iso-octane does not lead to theaverage of the laminar flame speed of ethanol and n-heptane orethanol and isooctane. Similar trends were also presented byBroustail et al. [10,11] for different conditions. Saisirirat et al. [12]conducted jet-stirred reactor (JSR) experiments for the ethanol/n-heptane (50/50, mole %) fuel mixture oxidation and found thataddition of ethanol in the fuel reduces the fuel consumption rates.For the auto-ignition characteristic studies of ethanol/iso-octaneblends, Bogin and co-workers [13] used the Ignition QualityTester (IQT) to investigate the effect of ethanol addition on theignition delay behavior of iso-octane at low temperatures withemphasize on the negative temperature (NTC) regime. Theyshowed that if the ethanol is more than 50%, NTC behavior van-ishes. Recently, Barraza-Botet and Wooldridge [14] used their rapidcompression facility (RCF) and gas chromatography (GC) to mea-sure the ignition delay time (IDT) of the blends and to identify theintermediate species formed during the ignition delay periods, andtheir results indicate that the ethanol and iso-octane are consumedin a parallel reaction pathway and have no significant fuel-to-fuelchemistry interactions. Li et al. [15] investigated the auto-ignitiontemperature and IDT of n-heptane/ethanol in a constant volumecombustion bomb.

Besides straight and branched chain alkanes, cyclic alkanes areanother component in jet fuels, gasoline, and diesel with a massfraction of around 20%, 10% and 40%, respectively [16e18]. Meth-ylcyclohexane (MCH) is selected as the cyclic alkane representativein transportation fuel surrogates [19,20]. There have been severalkinetic modeling efforts on MCH oxidation chemistry. Orme et al.[21] proposed a detailed high temperature mechanism for theoxidation of MCH and validated this kinetic scheme against IDTsobtained in their shock tube experiments. Pitz et al. [22] developeda low temperature kinetic model based on that of Orme et al. [21]by adding the low temperature reaction pathways of MCH oxida-tion. Subsequently, this model was updated respectively by Nar-ayanaswawy et al. [23] and Weber et al. [24] to get betteragreements with more experimental data [25]. Recently, Bissoo-nauth et al. [26] built a model of MCH based on the n-heptaneoxidation mechanism [29] with incorporated the high temperaturechemistry of MCH developed by Wang et al. [21] and the low

Fig. 1. Schematic of

temperature chemistry developed byWeber et al. [25] and referredto the low temperature classes for alkanes proposed by Sarathyet al. [30]. Because the model prediction for MCH oxidation in jetstirred reactor (JSR) and IDT in rapid compression machine shows ahigher reactivity than the experimental measurements, they pro-posed that additional data of MCH especially at low temperaturesare necessary for further model improvement.

As MCH is an important component in the transportation fuelsurrogate formulation, and the effect of ethanol addition into MCHon ignition characteristics has not been reported yet. The

the XJTU-RCM.

Page 3: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

Y. Liu et al. / Energy 177 (2019) 465e475 467

motivations of this work are as follows. Firstly, we will report newignition data for blends of ethanol and MCH obtained from a rapidcompression machine. These experimental data provides intuitiveinsights into the effect of temperature, pressure, equivalence ratioand ethanol blending ratio on IDT. Besides, these data will be usedto validate the newest mechanism developed by Bissoonauth et al.[26]. In addition, the intermediate species and the consumptionpath during ignition process will be scrutinized with the mecha-nism. Through this method, key species and reactions will beidentified to clarify the negative temperature coefficient (NTC)behavior and the effect of NTC regime shift due to ethanol chem-istry. Finally, because the effect of ethanol blending on the lowtemperature ignition behaviors of the C7 hydrocarbons withdifferent fuel structures will be compared to see if there is anygeneralized effect of ethanol chemistry. This study will be helpfulfor understanding the auto-ignition kinetics of binary fuels thatcontain ethanol and their possible application in engines.

2. Experimental apparatus and procedures

2.1. Rapid compression machine

IDTs can be determined by several types of devices, amongwhich rapid compression machine (RCM) are widely adopted forlow-to-intermediate temperatures measurements. In this study, apneumatically driven and hydraulically stopped rapid compressionmachine was used. The schematic of this RCM is shown in Fig. 1. Ahigh pressure gas tank is charged with compressed air to provide0.1e0.5MPa driving gas. The hydraulic chamber is filled with high

Fig. 3. Comparisons of the measured and simulated pressure traces for MCH/O2/N2/Armixtures at 20 bar. (a) f¼ 0.5, Tc¼ 722 K; (b) f¼ 1.0, Tc¼ 840 K.

pressure oil to hold the pistons at their resting positions before thecompression. To restrain the in-cylinder roll-up vortices and ensurethe homogeneous temperature distribution in the combustionchamber, a creviced piston like other RCMs [27,28] is adopted. Thecompression chamber of this RCM has a bore of 50.8mm and astroke length of 332mm. By adjusting the length of the reactionchamber from 10 to 100mm with different filling plates, thecompression ratio can vary from 4 to 15, so that the compressedtemperature at the end of compression (EOC) is changeable. Whenthe computer releases the signal of opening the oil discharge valve,the high pressure air drives the piston to compress the gas phasemixture. The chamber pressure is measured by the pressuretransducer (Kistler 6125C) and the charge amplifier (5018A), and itis recorded by a National Instrument USB-6361 data collecting cardwhich runs at a frequency of 100 kHz. Further details and thevalidation of this RCM are available in Ref. [29].

2.2. Definition and reproducibility of measured ignition delay times

Pressure and temperature at the EOC (Pc and Tc) are used todefine the thermodynamic conditions for IDT determination. Sometypical pressure-time (p-t) traces of MCH are shown in Fig. 2. Tc iscalculated according to the adiabatic compression relation, lnðPcPiÞ ¼R TcTi

gg�1

dTT , in which Ti and Pi are the initial pressure and tempera-

ture before the compression starts, and g is a temperature depen-dentmixture specific heat ratio. Pi and the length of the combustionchamber are altered to make the comparison of different mixtures

Fig. 4. Comparison between measured and simulated IDTs for neat MCH at Pc¼ 20 bar:(a) Total IDTs; (b) First-stage IDTs.

Page 4: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

Y. Liu et al. / Energy 177 (2019) 465e475468

is possible under the same pressure condition. And the uncertaintyin Tc is estimated to be less than 10 K by using the independentparameters methodology [30,31]. All experiments have beenrepeated with enough times and good repeatability was achieved.Fig. 1 shows the typical pressure evolution histories with threerepeating compressions at identical test conditions. The IDT at thiscondition is determined as the time interval between the EOC andthe time of peak pressure rise rate. It is noted that the pressureslightly falls down after EOC and this indicates that there are certainextent of heat loss from the core gases to the wall of the reactionchamber after the piston is locked at EOC. To account for these heatlosses in zero-dimensional simulations of the ignition process, non-reactive pressure traces are also collected by replacing the oxygencontent with nitrogen, as shown by the dash line in Fig. 2. Theinitial conditions of reactive and non-reactive tests are identical, sothe non-reactive pressure trace follows the reactive pressure trace.After the EOC, the non-reactive trace drops faster than the reactivetrace as no reactions happened to release heat.

2.3. Preparation of mixture and experimental conditions

The mixtures were prepared in a stainless-steel cylinder with avolume of 20 L, which contains sufficient test mixture to minimizethe uncertainty in mixture preparation. Since MCH and ethanol areliquid fuels at room temperature, they were injected via an injec-tion port on top of the cylinder using a gas-tight syringe. Purities ofnitrogen, argon and oxygen used in this study are higher than

Fig. 5. Comparison between experimental data and simulation results for blendmixtures at f¼ 0.5: (a) Total IDTs; (b) First-stage IDTs.

99.999%. MCH and ethanol were purchased from Energy Chemical(Shanghai, China) with purities of more than 99% and 99.5% byvolume, respectively. No further purification was carried out in thisstudy. Different fuel mixtures are prepared according to the partialpressure of ethanol and MCH and E25 represents that the fuelcontains 25% ethanol and 75% MCH in mole fraction. The equiva-lence ratio is calculated with considering the oxygen contained inethanol. The experimental region of Pc and Tc are listed in Table 1.For each mixture, nitrogen and argon are used as the buffer gaswith a 85% mole fraction and the specific components and pro-portions can be found in Supplementary materials.

3. Results and discussion

3.1. Mechanisms validation against new RCM data of pure MCH

Numerical simulation of ignition process is conducted usingChemkin-Pro 15131 software and the volume history tabulation isadopted, where the volume history of non-reactive mixtures istransformed from the pressure trace at corresponding condition toaccount for facility effect during the compression. And the kineticmodel of Bissoonauth et al. [26] is used for the simulation, thus thesimulated pressure history from compression to ignition is ob-tained. Typical simulation cases of single-stage and two-stageignition are shown in Fig. 3 for further illustration. The non-reactive pressure traces overlap with the reactive pressure tracesbefore the EOC at the same condition, indicating that the heatrelease caused by reactions before EOC is negligible. Similarly, the

Fig. 6. Comparison between experimental data and simulation results for blendmixtures at f¼ 1.0: (a) Total IDTs; (b) First-stage IDTs.

Page 5: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

Y. Liu et al. / Energy 177 (2019) 465e475 469

moments of the maximum pressure rise in simulated pressuretraces are defined as the ignition onsets. Thus, the simulated IDTsare the interval time between the EOC and the ignition onsets. InFig. 3a, the measured and simulated pressure traces both show atwo-stage pressure rise behavior, indicating the typical two stageignition for this mixture. The measured first stage IDT and total IDTwere respectively 15.4 and 19.6ms, while the simulated first stageIDT and total IDT were 10.5 and 26.0ms. For higher temperaturecase as shown in Fig. 3b, the two-stage ignition phenomenon dis-appears and the prediction of total IDT is longer than the measuredvalue.

The IDTs of MCH/O2/N2/Ar mixtures at f¼ 0.5 and 1.0,Pc¼ 20 bar were measured over the Tc range between 690 and950 K. The total IDTs are plotted against 1000/Tc as shown in Fig. 4a.The total IDT shows a clear NTC behavior, and themixture reactivityincreases with the equivalence ratio increases. This tendency isconsistent with the observation by Weber et al. [24]. We note thatin this study, the inert concentration is fixed and the fuel concen-tration is fixed in the experiments conducted by Weber et al. [24].The NTC regime, within which the total IDT increases withincreasing Tc, is bounded by two turnover temperatures, i.e., Tupperand Tlower as defined in Ref.s [32,33]. In this NTC regime, increasingtemperature actually reduces the overall reactivity. For this pureMCH fuel, the first-stage IDTs are also presented, as shown inFig. 4b. With the increase of f, the first-stage IDT decreases, thoughthis effect becomes progressively weakened as the temperature is

Fig. 7. Comparison between experimental data and simulation results for the stoi-chiometric mixture E50 at different pressures: (a) Total IDTs; (b) First-stage IDTs.

reduced. the numerical prediction with the model of Bissoonauthet al. [26] shows a shorter first-stage IDT and longer total IDT,compared with the experimental data. The max deviation happensat the temperature regime where the first-stage IDTs are relativelyshort. Generally, the model well captures the two-stage ignitionbehavior with reasonable agreement between its prediction of thetotal IDT, though the model slightly under-estimate the first stageIDT.

3.2. The effect of ethanol blending ratio

To investigate the effects of blending ethanol to MCH, thepressure traces of different MCH/ethanol mixtures at Pc¼ 20 barand Tc from 680 to 930 K are measured. The comparison betweenthe model prediction and experimental data for f¼ 0.5 and 1.0 areshown respectively in Fig. 5 and Fig. 6. Symbols represent theexperimental data while lines represent the simulation, error barsare omitted for concision, and different ethanol blend conditionsare performed in different colors. It is noted that blending ethanolintoMCH prolongs the total IDTs of themixture significantly at bothfuel lean and stoichiometric conditions. This is reasonable becausethe ignitability of the binary blending mainly depends on thereactive fuel with low octane number [15]. The more reactivespecies ignites first and releases heat which results in theincreasing of temperature and pressure, and then accelerates theignition of the less reactive species. As such, for high ethanolblending ratio mixture like E75, it cannot ignite at lower temper-atures thus neither experimental nor simulation data are shown forthese fuel blends. Literature indicate that there is no NTC behaviorfor pure ethanol [34,35] while pure MCH has NTC behavior [24,25],which is consistent with the present observation. In addition, we

Fig. 8. Computed temperature and species concentrations during ignition process forstoichiometric mixture E50 at Pi¼ 20 bar. (a)Ti¼ 725 K; (b) Ti¼ 875 K.

Page 6: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

Fig. 9. Computed temperature histories for different mixtures at Ti¼ 725 K andPi¼ 20 bar.

Y. Liu et al. / Energy 177 (2019) 465e475470

observed that the NTC regime shifts to lower temperatures andnarrows down alongwith the increase of ethanol content in Figs. 5aand 6a. When the ethanol content reaches a certain extent, the NTCregime disappears, as shown by the data of E75 in Fig. 5a and E100in Fig. 6a. For the model simulation, the simulated total IDTs areslightly longer than the measured data around Tlower and slightlyshorter than the measured data around Tupper. We also note that thefirst-stage ignition phenomenon disappears when the temperaturereaches to a certain extent as in Figs. 5b and 6b, and that temper-ature limit becomes lower as ethanol fraction increases. For thefirst-stage ignition simulation, the model underestimated the first-stage IDT values and failed to precisely predict whether the first-stage ignition will happen. For example, the experiments showtwo-stage ignition phenomenon whereas the simulations showsonly one stage ignition for the mixture E75. As such further im-provements for this mechanism will be needed in the future.

3.3. The effect of pressure

Fig. 7 shows the comparison of IDTs for the stoichiometricmixture of E50 at Pc¼ 10, 20, and 30 bar. Higher Pc accelerates theignition significantly: the total IDT at Pc¼ 10 bar is almost an orderof magnitude higher than that at 30 bar. This is straightforwardbecause higher temperature and pressure accelerate the speed ofmolecules and number density of themolecules thus enhancing thecollision frequency which results in higher rate of chemical reac-tion. It is noted that the mixture can ignites at lower temperaturesbut fails to ignite when the temperature is higher than 790 K atPc¼ 10 bar. It could be explained from the tendency that the totalIDT increases with Tc as far as the temperature is in the NTC regime.As a result, no ignition is observed in experiments for temperaturehigher than 790 K. In addition, the temperature limit where thefirst-stage ignition disappears is shifted to high temperatures as Pcincreases, as shown in Fig. 7b. For the mixture E50, good agree-ments between measured and simulated data for the total IDTs,while simulations for the first-stage IDTs are still slightly shorterthan the experimental data.

4. Chemical kinetic analysis of binary fuel ignition

4.1. Species profiles of binary mixture

To understand the effect of ethanol addition on the ignitionkinetics of MCH, the model of Bissoonauth et al. [26] is used tocompute the species and temperature evolution during the ignitionprocess of the binary fuel mixtures. The species profiles for stoi-chiometric mixture E50 at initial pressure (Pi) of 20 bar, and initialtemperature (Ti) of 725 and 875 K are computed with Chemkin-Pro15131. The species selected for kinetic analysis include the fuels, OHradical and H2O2 radicals as they are generally the ignition in-dicators [15,36]. Fig. 8a and b shows the typical two-stage ignitionand single-stage ignition phenomenon, respectively. The initialmole fraction of MCH and ethanol are the same in mixture E50, butthe mole fraction of MCH drops apparently faster than ethanolwhich indicates that MCH plays a leading role in competition ofsmall radicals such as O, H and OH radicals which consumes the fuelthrough H-abstraction reactions. The first-stage ignition in Fig. 8ahappens when the consumption rate of MCH and ethanol reach thefirst peak, the H2O2 molecule also reaches the first maximum for-mation rate at this time. It is found that the first-stage ignitionoccurs at a relative lower system temperature compared to thesecond-stage ignition, and the second-stage ignition starts at themoment when the OH radical concentration begins to soar. Duringthe first-stage ignition, the OH radical concentration changes littlebut the model fraction of H2O2 molecule increase rapidly, and the

H2O2 starts to dissociate rapidly only until the system temperatureapproaches approximately 1000 K [37]. The produced OH radicalswill consume the remaining fuels by H-abstraction reactions whichfurther elevate the system temperature significantly throughreleasing the heat. That is to say, the temperature limit of H2O2decomposition can be regarded as a threshold for the ignitionprocess, so the total IDT is actually the time spent for the system toreach that temperature threshold. To validate this assumption, thetemperature histories of different mixtures at identical conditionswere conducted as shown in Fig. 9. Neat MCH (E0) ignites first sinceits first-stage ignition occurs first and provides the largest tem-perature increase, and the heat release during the first-stage igni-tion decreases as more ethanol is blended in. The lowertemperature rise in the first-stage ignition directly causes thelonger second-stage IDT. All the mixtures ignite when the systemtemperature reaches about 1000 K under which condition H2O2begins to decompose rapidly, and this indicates that the finalignition is independent of other factors but only related to thesystem temperature.

4.2. Reaction fluxes of neat MCH and binary fuels

With the basic understanding of ignition process, we then try toelucidate the NTC regime shift phenomenon by analyzing the re-action flux during ignition. This analysis is conducted using aconstant-volume and adiabatic type simulation for mixture E0 andE50 at Pi¼ 20 bar, and three different Ti (725, 800, and 875 K). Sincethe NTC behavior occurs for MCH species, we mainly focus on theconsumption path of MCH. The path analysis presented in Fig. 10 isan integrated analysis where the rate of production of each speciesby each reaction has been integrated with respect to time up to 20%MCH consumption, so the percentages represent the percent of thegiven reactant that is consumed to form the given product by allreactions that can form that particular product. Note that not all thepossible reaction pathways are shown in Fig. 10 for simplicity.

For the primary step of fuel consumption, Narayanaswamy et al.[23] presented the reaction fluxes at Ti¼ 1160 K and found thatMCH mainly undergoes unimolecular decomposition reactions.However, this is not observed in the present analysis. This indicatesthat the favored consumption channel in low temperatures(<1000 K) is changed from unimolecular decomposition reactionsto H-abstraction reactions. There are five different sites for H-atomabstraction due to the symmetry of MCH, thus five methyl-cyclohexyl radicals are formed through bimolecular reactions withOH and HO2.

Page 7: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

Y. Liu et al. / Energy 177 (2019) 465e475 471

Because nearly all the fuel radicals are oxidized to the alkyl-peroxy radical (ROO), and there are three main reaction classes thatparticipate in the subsequent reactions of ROO radicals. The firstclass is the intramolecular isomerization to form hydroperoxyalkylradical (QOOH). The second class is the elimination reaction to formhydroperoxyl (HO2) and methylcyclohexene molecule, and thethird class is the H-abstraction by ROO to form ROOH. TakeMCH2OO as a specific example. For the reactions among MCH2OOradicals to produce MCH2OJ, they consume 3.9% of MCH2OO at700 K and less than 0.1% for the other temperatures, so thispathway is not shown in Fig. 9. The majority of MCH2OJ radical isproduced through the third class reactions. Specifically, H-abstraction reactions from HO2 by MCH2OO produce MCH2OOH

Fig. 10. Path analysis of MCH consumption for mixture E0 (plain text) and mixture E50 (boldpossible reaction pathways are shown for each species. (For interpretation of the references

and an oxygen molecule, the former of which then decomposes toMCH2OJ and OH radical. The MCH2OJ subsequently undergoes ringopen reactions and b-scission to produce small olefins and alde-hydes with no more than 5 carbon atoms. This channel increasesthe mixture reactivity because OH radical is produced. However, itscontribution for MCH2OO consumption quickly decreases from21.8% to less than 5% as the Ti increases from 700 K to 800 K. Thisreveals that the formation of ROOH becomes substantially lessimportant and the competition of MCH2OO consumption mainlyhappens between the first and second class reactions at low tem-peratures (<1000 K). The first class reactions produce QOOH whichthen faces second oxygen addition or decomposition. The formerpathway produces hydroperoxyalkylperoxy (OOQOOH) radical

text) at Pi¼ 20 bar, Ti¼ 725 K (blue), 800 K (black), and 875 K (red). Note that not all theto color in this figure legend, the reader is referred to the Web version of this article.)

Page 8: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

Y. Liu et al. / Energy 177 (2019) 465e475472

which is then consumed by concerted elimination to olefinic hy-droperoxides (C7H12O2) and HO2 radical or decomposition toketohydroperoxides (C7H12O3) and OH radical. This is a typicalchain branching reaction which is more favored at lower temper-atures and leads to the two-stage ignition phenomenon. Thecompeting channel, decomposition of QOOH to form a heptenoneand OH radical, becomes more and more important as Ti increases.On the other hand, the second class reaction also contributesincreased consumption for ROO as Ti increases, and the increase inproduction of methylcyclohexene and HO2 plays a crucial role inthe NTC regime. The reason is that HO2 attacks H-atom from thefuel to form H2O2 which is relatively stable in the low temperatureregime(<1000 K) and this channel prolongs the overall ignition asdescribed above.

The OOQOOH formation channel plays the dominating role inROO consumption at low temperatures (<725 K), and the large heatrelease results in first-stage ignition phenomenon behavior. Withthe increase of Ti, the ROO consumption of first and third class re-action reduces and the second class reaction consumes more ROOto compensate, and the HO2 elimination of ROOH takes the leadingrole compared to the OOQOOH formation channel at a certaintemperature which is named the “ceiling temperature” by Bensonet al. [38]. When Ti is higher than this ceiling temperature, the alkylfuel radical mainly undergoes b-decomposition to produce olefinand HO2 which restrains the reactivity and causes the disappear-ance of two-stage ignition behavior and NTC regime of total IDT.Though more H2O2 are produced as Ti increases, the gap betweenthe system initial temperature and H2O2 decomposition tempera-ture becomes narrowed. Finally, when Ti exceeds a certain value,

Fig. 11. Effect of ethanol addition on the radical mole fraction at Ti¼ 725 K, Pi¼ 20 bar,and f¼ 1.0. (a) ROO radicals; (b) QOOH radicals.

H2O2 decomposes immediately as they are produced. As a conse-quence, the overall IDT decreases again as the Ti rises.

The original reason of NTC behavior is all related to the forma-tion of ROO and QOOH, and these species are not formed in theethanol oxidation [34].This is the primary reason that ethanol doesnot exhibit NTC behavior. It is also found in Fig. 10 that ethanolblending changes the consumption of MCH slightly. However, inthis study, the mole fraction of dilution is fixed at 85%, so theethanol blending results the decrease of MCH mole fraction. Fig. 11

Fig. 12. Total IDT as a function of the temperature for stoichiometric fuel/ethanol/O2/Armixtures under various ethanol blending ratios. (a) MCH; (b) n-heptane; (c) 1-heptene.

Page 9: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

Y. Liu et al. / Energy 177 (2019) 465e475 473

shows the computed mole fraction time-histories of ROO andQOOH for different blending ratio at the same initial condition. TheROO radicals include MCH1OO, MCH2OO, MCH3OO, MCH4OO, andCHXCH2OO, while the QOOH radicals include 11 species such asMCH1QJ2, MCH1QJ3, MCH1QJ4, MCHXQJ3, etc. It is noted that themole fraction time-histories of both ROO and QOOH decreasesrapidly as the ethanol is presented in the system, and the reductionof their peak value indicates fewer oxygen addition reactionshappen during low temperatures. This reveals that the NTCbehavior becomes less apparent as ethanol is added, which isconsistent with the experimental observation as in Figs. 5 and 6. Forthe effect of pressure, the overall ignition threshold reaction: H2O2(þM) ¼ 2OH(þM) is a typical pressure-dependent reaction whoserate increases with increasing pressure, this results the NTC regimeshifts to higher temperature as the pressure increases [32].

4.3. Effect of ethanol blending ratio on the turnover temperatures

Ji et al. [32,33] quantified the temperature range and slope of theNTC regime, and found that the total IDTs shows Arrhenius liketemperature dependence for different alkanes with the turnovertemperatures under various pressures for different alkanes. Wenow investigate whether this relationship still holds underdifferent ethanol blending ratios. MCH, n-heptane, and 1-hepteneare selected because all of them perform NTC behavior in previ-ous experiments [24,30,39] and are included in the mechanism of

Fig. 13. Computed total IDTs at turnover temperatures for stoichiometric fuel/ethanol/O2/Ar mixtures under various ethanol blending ratios. (a) total IDTs at Tupper; (b) totalIDTs at Tlower.

Bissoonauth et al. [26]. The computed total IDTs as a function oftemperature for a stoichiometric mixture of fuel/ethanol/O2/Arwith different ethanol blending ratios were shown in Fig. 12. Theconcentration of fuel is fixed at 1.0% and Pi is fixed at 20 bar in allthemixtures. It is seen that for each ethanol blending ratio less than80%, the total IDTs decrease with increasing the initial temperatureexcept the NTC regime. Both the Tlower and Tupper shift to lowervalues as ethanol blending ratio is increased, and the NTC regimeshrinks accordingly. It is noted that the total IDTs does not showNTC behavior when the ethanol blending ratio reaches 80%. Andthe IDTs at turnover temperatures, marked by the triangle symbolsin Fig. 12, show Arrhenius like temperature dependence whenethanol blending ratios is no more than 60%.

We next note that the total IDTs exhibit special dependence onthe ethanol blending ratio at the turnover temperatures. Fig. 13suggests a parabolic curve fitting, namely, IDT¼aRE

2þbRE þ c,where RE is the ethanol blending ratio. To further investigate theeffect of ethanol blending ratio on total IDTs at turnover tempera-tures, we normalized the total IDTs according to the total IDTwithout ethanol blending as shown in Fig. 14. It is noted that thevariation of normalized total IDTs at turnover temperatures is verysimilar for MCH and n-heptane, especially at upper turnover tem-perature. And the total IDTs of 1-heptene increase slower than theother two fuels at turnover temperatures with ethanol blending.

We finally investigate the relationship between turnover

Fig. 14. Normalized total IDTs at turnover temperatures for stoichiometric fuel/ethanol/O2/Ar mixtures under various ethanol blending ratios. (a) normalized IDTs atTupper; (b) normalized IDTs at Tlower.

Page 10: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

Y. Liu et al. / Energy 177 (2019) 465e475474

temperature and ethanol blending ratio as shown in Fig. 15. It isnoted that as the ethanol blending ratio increases to 60%, the Tupperand Tlower of three fuels all decrease about 70 K and 30 K, respec-tively. Previous works [14,15] indicate that there are no fuel-to-fuelinteractions (ethanol to n-heptane or iso-octane), and the reducedreactivity and the shrink and shift of NTC regime by ethanol addi-tion is attributed to the decreased n-heptane and iso-octane con-centration, which leads to reduced production of ROO and QOOH asethanol fraction increases. The similar ethanol addition effect onIDT and NTC regime behavior of MCH and n-heptene with thatreported in Refs. [14,15] indicate that for a general binary ethanol/hydrocarbon fuel system, it is unlikely to have fuel to fuel interac-tion, and the presence of ethanol just simply reduces the concen-tration of the hydrocarbon.

5. Conclusions

In this study, experimental measurements of low temperatureIDT of MCH and ethanol binary fuels have been conducted overwide range of test conditions. Measurements show that pure MCHexhibits typical two-stage ignition and the negative temperaturebehavior (NTC). With the increase of ethanol blending ratio, theoverall reactivity of the fuel blends decreases. In addition, the NTCregimewhich is bounded by two turnover temperatures (Tupper andTlower) is shifted to lower temperatures and the difference betweenthe two turnover temperatures decreases. These new IDT datawerethen used to validate the most updated mechanism of MCH [26].

Fig. 15. Turnover temperatures for stoichiometric fuel/ethanol/O2/Ar mixtures undervarious ethanol blending ratios. (a) Tupper; (b) Tlower.

Good agreements were observed with the measured and simulatedtotal IDTs, while the model still show an underestimated the first-stage IDT.

To understand the kinetic effect of ethanol addition on the IDTsand the NTC behavior, species evolution process and reaction fluxanalysis were conducted. The results showed that the final ignitionoccurs only when the system temperature reaches the decompo-sition temperature of H2O2, and the nature of NTC behavior is thecompetition between chain branching and chain determinationreactions under different temperatures. The main chain branchingreactions and chain determination reactions are the oxygen addi-tion reactions and the HO2 elimination reactions, respectively.Although no significant change on MCH consumption path wasobserved as ethanol is blended in, the mole fraction of ROO andQOOH decrease rapidly due to the reduction of MCH initial con-centration. Therefore, the MCH/ethanol blend acts as a super-position of the reaction chemistry of the two individuals, andconsequently the NTC behavior becomes less apparent as ethanol isblended in.

Further comparison among the ignition kinetics of the differentbinary fuels including, MCH, n-heptane, and 1-heptene have beenconducted to access the different structure effect. The NTC regimeshrink and shift phenomenon with increasing ethanol blendingratio are observed for all the three binary fuels investigated. Inaddition, both total IDTs and turnover temperatures show similarvariation tendency as the ethanol blending ratio increases. Thissimilarity indicates that for a general binary ethanol/hydrocarbonsystem, fuel-to-fuel interactions are absent and the presence ofethanol just simply reduces the hydrocarbon concentration. How-ever, further experimental examination of the intermediate speciesduring the oxidation of binary ethanol/hydrocarbon fuels is of meritfor kinetic mechanism validation.

Acknowledgments

This work is supported by the National Natural Science Foun-dation of China (51722603, 91541107, and 91441203), ScienceChallenging Program (TZ201601), the Fundamental Research Fundsfor the Central Universities, and Key Laboratory of High Efficiencyand Low Emission Engine Technology, Ministry of Industry andInformation Technology of the People's Republic of China, BeijingInstitute of Technology(2017CX02015).

Appendix A. Supplementary data

Supplementary data to this article can be found online athttps://doi.org/10.1016/j.energy.2019.04.132.

References

[1] Hansen AC, Zhang Q, Lyne PWL. Ethanolediesel fuel blends ee a review.Bioresour Technol 2005;96(3):277e85.

[2] Niven RK. Ethanol in gasoline: environmental impacts and sustainability re-view article. Renew Sustain Energy Rev 2005;9(6):535e55.

[3] Matti Maricq M. Soot formation in ethanol/gasoline fuel blend diffusionflames. Combust Flame 2012;159(1):170e80.

[4] Khosousi A, Liu F, Dworkin SB, Eaves NA, Thomson MJ, He X, et al. Experi-mental and numerical study of soot formation in laminar coflow diffusionflames of gasoline/ethanol blends. Combust Flame 2015;162(10):3925e33.

[5] Bierkandt T, Kasper T, Akyildiz E, Lucassen A, Oßwald P, K€ohler M, et al. Flamestructure of a low-pressure laminar premixed and lightly sooting acetyleneflame and the effect of ethanol addition. Proc Combust Inst 2015;35(1):803e11.

[6] Lü X, Ji L, Zu L, Hou Y, Huang C, Huang Z. Experimental study and chemicalanalysis of n-heptane homogeneous charge compression ignition combustionwith port injection of reaction inhibitors. Combust Flame 2007;149(3):261e70.

[7] Mack JH, Aceves SM, Dibble RW. Demonstrating direct use of wet ethanol in ahomogeneous charge compression ignition (HCCI) engine. Energy 2009;34(6):

Page 11: Low temperature auto-ignition characteristics of ...gr.xjtu.edu.cn/upload/22045/Low+temperature+auto...Low temperature auto-ignition characteristics of methylcyclohexane/ ethanol blend

Y. Liu et al. / Energy 177 (2019) 465e475 475

782e7.[8] Isık MZ, Aydın H. Analysis of ethanol RCCI application with safflower biodiesel

blends in a high load diesel power generator. Fuel 2016;184:248e60.[9] Lipzig JPJV, Nilsson EJK, Goey LPHD, Konnov AA. Laminar burning velocities of

n -heptane, iso-octane, ethanol and their binary and tertiary mixtures. Fuel2011;90(8):2773e81.

[10] Broustail G, Seers P, Halter F, Mor�eac G, Mounaim-Rousselle C. Experimentaldetermination of laminar burning velocity for butanol and ethanol iso-octaneblends. Fuel 2011;90(1):1e6.

[11] Broustail G, Halter F, Seers P, Mor�eac G, Mounaïm-Rousselle C. Experimentaldetermination of laminar burning velocity for butanol/iso-octane and ethanol/iso-octane blends for different initial pressures. Fuel 2013;106:310e7.

[12] Saisirirat P, Togb�e C, Chanchaona S, Foucher F, Mounaim-Rousselle C,Dagaut P. Auto-ignition and combustion characteristics in HCCI and JSR using1-butanol/n-heptane and ethanol/n-heptane blends. Proc Combust Inst2011;33(2):3007e14.

[13] Bogin GE, Luecke J, Ratcliff MA, Osecky E, Zigler BT. Effects of iso-octane/ethanol blend ratios on the observance of negative temperature coefficientbehavior within the Ignition Quality Tester. Fuel 2016;186:82e90.

[14] Barraza-Botet CL, Wooldridge MS. Combustion chemistry of iso-octane/ethanol blends: effects on ignition and reaction pathways. Combust Flame2018;188:324e36.

[15] Li R, Liu Z, Han Y, Tan M, Xu Y, Tian J, et al. Experimental and kinetic modelingstudy of autoignition characteristics of n-heptane/ethanol by constant volumebomb and detail reaction mechanism. Energy Fuels 2017;31(12):13610e26.

[16] Briker Y, Ring Z, Iacchelli A, McLean N, Rahimi PM, Fairbridge C, et al. Dieselfuel analysis by GC�FIMS: aromatics, n-paraffins, and isoparaffins. EnergyFuel 2001;15(1):23e37.

[17] Pitz WJ, Mueller CJ. Recent progress in the development of diesel surrogatefuels. Prog Energy Combust Sci 2011;37(3):330e50.

[18] Sarathy SM, Farooq A, Kalghatgi GT. Recent progress in gasoline surrogatefuels. Prog Energy Combust Sci 2018;65:67e108.

[19] Humer S, Frassoldati A, Granata S, Faravelli T, Ranzi E, Seiser R, et al. Exper-imental and kinetic modeling study of combustion of JP-8, its surrogates andreference components in laminar nonpremixed flows. Proc Combust Inst2007;31(1):393e400.

[20] Violi A, Yan S, Eddings EG, Sarofim AF, Granata S, Faravelli T, et al. Experi-mental formulation and kinetic model for JP-8 surrogate mixtures. CombustSci Technol 2002;174(11e12):399e417.

[21] Orme JP, Curran HJ, Simmie JM. Experimental and modeling study of methylcyclohexane pyrolysis and oxidation. J Phys Chem A 2006;110(1):114.

[22] Pitz WJ, Naik CV, Mhaoldúin TN, Westbrook CK, Curran HJ, Orme JP, et al.Modeling and experimental investigation of methylcyclohexane ignition in arapid compression machine. Proc Combust Inst 2007;31(1):267e75.

[23] Narayanaswamy K, Pitsch H, Pepiot P. A chemical mechanism for low to hightemperature oxidation of methylcyclohexane as a component of trans-portation fuel surrogates. Combust Flame 2015;162(4):1193e213.

[24] Weber BW, Pitz WJ, Mehl M, Silke EJ, Davis AC, Sung CJ. Experiments and

modeling of the autoignition of methylcyclohexane at high pressure. CombustFlame 2014;161(8):1972e83.

[25] Mittal G, Sung CJ. Autoignition of methylcyclohexane at elevated pressures.Combust Flame 2009;156(9):1852e5.

[26] Bissoonauth T, Wang Z, Mohamed SY, Wang J-y, Chen B, Rodriguez A, et al.Methylcyclohexane pyrolysis and oxidation in a jet-stirred reactor. ProcCombust Inst 2019;37(1):409e17.

[27] Brett L, Macnamara J, Musch P, Simmie JM. Simulation of methane auto-ignition in a rapid compression machine with creviced pistons. CombustFlame 2001;124(1):326e9.

[28] Mittal G, Raju MP, Sung CJ. Computational fluid dynamics modeling ofhydrogen ignition in a rapid compression machine. Combust Flame2008;155(3):417e28.

[29] Liu Y, Tang C, Wu Y, Yang M, Huang Z. Low temperature ignition delay timesmeasurements of 1,3,5-trimethylbenzene by rapid compression machine. Fuel2019;241:637e45.

[30] Wu Y, Liu Y, Tang C, Huang Z. Ignition delay times measurement and kineticmodeling studies of 1-heptene, 2-heptene and n-heptane at low to interme-diate temperatures by using a rapid compression machine. Combust Flame2018;197:30e40.

[31] Xu N, Wu Y, Tang C, Zhang P, He X, Wang Z, et al. Experimental study of 2,5-dimethylfuran and 2-methylfuran in a rapid compression machine: compar-ison of the ignition delay times and reactivity at low to intermediate tem-perature. Combust Flame 2016;168:216e27.

[32] Ji W, Zhao P, He T, He X, Farooq A, Law CK. On the controlling mechanism ofthe upper turnover states in the NTC regime. Combust Flame 2016;164:294e302.

[33] Ji W, Zhao P, Zhang P, Ren Z, He X, Law CK. On the crossover temperature andlower turnover state in the NTC regime. Proc Combust Inst 2017;36(1):343e53.

[34] Mittal G, Burke SM, Davies VA, Parajuli B, Metcalfe WK, Curran HJ. Auto-ignition of ethanol in a rapid compression machine. Combust Flame2014;161(5):1164e71.

[35] Cancino LR, Fikri M, Oliveira AAM, Schulz C. Measurement and chemical ki-netics modeling of shock-induced ignition of ethanol�air mixtures. EnergyFuels 2010;24(5):2830e40.

[36] Zheng Y, Yong Q, Xin Y, Yue W, Wang Y, Zhen H, et al. Autoignition of n-butanol/n-heptane blend fuels in a rapid compression machine under low-to-medium temperature ranges. Energy Fuels 2013;27(12):7800e8.

[37] Westbrook CK. Chemical kinetics of hydrocarbon ignition in practical com-bustion systems. Proc Combust Inst 2000;28(2):1563e77.

[38] Benson SW. The kinetics and thermochemistry of chemical oxidation withapplication to combustion and flames. Prog Energy Combust Sci 1981;7(2):125e34.

[39] Mehl M, Pitz WJ, Westbrook CK, Curran HJ. Kinetic modeling of gasolinesurrogate components and mixtures under engine conditions. Proc CombustInst 2011;33(1):193e200.


Recommended