Effects Of Heat Of Vaporization And Octane Sensitivity On .

2y ago
6 Views
2 Downloads
1.93 MB
14 Pages
Last View : 1m ago
Last Download : 3m ago
Upload by : Louie Bolen
Transcription

NREL/CP-5400-70443. Posted with permission. Presented at WCX18: SAE World Congress Experience, 10-12 April 2018, Detroit, Michigan.2018-01-0218Published 03 Apr 2018Effects of Heat of Vaporization and OctaneSensitivity on Knock-Limited Spark Ignition EnginePerformanceMatthew A. Ratcliff, Jonathan Burton, Petr Sindler, Earl Christensen, Lisa Fouts, andRobert L. McCormick National Renewable Energy LaboratoryCitation: Ratcliff, M.A., Burton, J., Sindler, P., Christensen, E. et al., “Effects of Heat of Vaporization and Octane Sensitivity onKnock-Limited Spark Ignition Engine Performance,” SAE Technical Paper 2018-01-0218, 2018, doi:10.4271/2018-01-0218.AbstractKnock-limited loads for a set of surrogate gasolines allhaving nominal 100 research octane number (RON),approximately 11 octane sensitivity (S), and a heat ofvaporization (HOV) range of 390 to 595 kJ/kg at 25 C wereinvestigated. A single-cylinder spark-ignition engine derivedfrom a General Motors Ecotec direct injection (DI) engine wasused to perform load sweeps at a fixed intake air temperature(IAT) of 50 C, as well as knock-limited load measurementsacross a range of IATs up to 90 C. Both DI and pre-vaporizedfuel (supplied by a fuel injector mounted far upstream of theintake valves and heated intake runner walls) experimentswere performed to separate the chemical and thermal effectsof the fuels’ knock resistance. The DI load sweeps at 50 Cintake air temperature showed no effect of HOV on the knocklimited performance. The data suggest that HOV acts as athermal contributor to S under the conditions studied.IntroductionThere is a significant opportunity to improve sparkignition (SI) engine efficiency through the combinedapplication of more efficient engine designs and operating strategies, and improved fuels [1]. The benefits ofimproved SI engine efficiency are enhanced national securitythrough reduction of dependence on petroleum imports [2],reduced costs to consumers [3], and protection of the environment [4]. In the United States and around the world, governments are requiring significant vehicle fuel economy improvements to achieve these benefits [5,6].The primary fuel related engine design and operationalstrategies that improve efficiency are increased compressionratio (CR), the combined effects of engine downsizing andturbocharging, operating the engine at lower speeds, cylinderdeactivation, exhaust gas recirculation (EGR), and utilizingdirect injection (DI) [1,7,8]. Increasing CR increases the thermodynamic efficiency of the engine [9], while also increasingthe temperature and pressure of the unburned fuel-air mixtureresulting in engine knock at high loads. Thus, while increasingCR can be used to improve part load efficiency, it is limitedby the knock resistance of the fuel. Engines exhibit lower 2018 SAE International; Ford Motor Company; National Renewable Energy Laboratory.Measurement of knock-limited loads from the IAT sweeps forDI at late combustion phasing showed that a 40 vol% ethanol(E40) blend provided additional knock resistance at thehighest temperatures, compared to a 20 vol% ethanol blendand hydrocarbon fuel with similar RON and S. Using the prevaporized fuel system, all the high S fuels produced nearlyidentical knock-limited loads at each temperature across therange of IATs studied. For these fuels RON ranged from 99.2to 101.1 and S ranged from 9.4 to 12.2, with E40 having thelowest RON and highest S. The higher knock-limited loads forE40 at the highest IATs examined were consistent with theslightly higher S for this fuel, and the lower engine operatingcondition K values arising from use of this fuel. The studyhighlights how fuel HOV can affect the temperature at intakevalve closing, and consequently the pressure-temperaturehistory of the end gas leading to more negative values of K,thereby enhancing the effect of S on knock resistance.efficiency when operated at light loads because of parasiticlosses. Downsizing, turbocharging, and DI are used to operateengines at higher load over a larger portion of the enginemap [10]. Similarly, operating at lower engine speeds or downspeeding requires higher load to achieve the same power.Downsizing and down-speeding also reduce friction in theengine resulting in improved efficiency. Cylinder deactivationis used at light loads to increase the load and hence efficiencyin the remaining cylinders [11]. All these strategies to operatethe engine at higher loads also result in higher temperatureand pressure of the unburned fuel-air mixture, and thus arelimited by engine knock.Spark-timing retard is applied at high load conditions tomitigate knock but at the expense of reduced efficiency dueto less ideal combustion phasing, reduced torque, andincreased exhaust temperature [12]. At very high loads fuelenrichment is also used to reduce knock and control exhausttemperature to prevent engine and catalyst damage - but atthe expense of high carbon monoxide and hydrocarbonemissions [13].The high-efficiency engine design and operation strategies described above can all be pursued more aggressively-that

Effects of Heat of Vaporization and Octane Sensitivityis, with less use of spark timing retard and fuel enrichment-ifthe engine is using a more highly knock resistant fuel. Forexample, an analysis of several studies suggests that anincrease of 1 CR unit requires a research octane number(RON) increase of approximately 5 units [14]. Fuel propertiesthat affect knock resistance include RON, motor octanenumber (MON), octane sensitivity (S RON – MON), andheat of vaporization (HOV). Modern engines, especiallydownsized boosted engines, operate under conditions whereincreased RON and increased S (or decreased MON) causeincreased knock resistance [15,16]. This effect has beendescribed in terms of octane index (OI)-a fuel’s actual resistance to autoignition [17]-where:OI RON - K * S(1)For downsized boosted engines K is negative at the mostknock-limited operating conditions, such that increasing Sincreases OI [18].While retaining stoichiometric conditions, EGR improvesefficiency by reducing pumping losses at light and intermediate loads and by reduced heat transfer because of lower peaktemperature [7]. The specific heat ratio Cp/Cv (or γ) alsoincreases with EGR which improves the efficiency of theconversion of heat into work. Also, Alger and coworkers haveshown that the lower in-cylinder temperature caused by EGRreduces engine knock and that each percentage increase inEGR is equivalent to an octane number increase of 0.5 [8].DI enables higher power density in downsized boostedengines primarily by reducing the fuel-air mixture temperature through evaporative cooling [19, 20, 21, 22, 23]. Thesestudies suggest that cooling for hydrocarbon fuels providesthe same effect as an increase of roughly 5 octane numberunits-adequate to allow increasing compression ratio by 1 unit[15]. Evaporative cooling also improves engine volumetricefficiency by reducing pumping losses at part load [19]. Ethanolexhibits a much higher HOV than hydrocarbon gasolines, andthus ethanol blends can produce a significant increase inevaporative cooling and measured fuel knock resistance [19,21, 22, 23, 24, 25, 26, 27]. Kasseris and Heywood presentexperiments and analysis showing that charge cooling canrange from roughly 15 C for E0 to 30 C at E50 and as high as50 C for E85 blends, and that these values are about 70% ofthe thermodynamic maximum possible cooling [22].An important study by Foong and coworkers revealedthe impact of evaporative cooling from ethanol blending onthe RON measured in the Cooperative Fuels Research (CFR)engine by ASTM D2699 [28]. These researchers found littleimpact of increased HOV on RON up to approximately 30 to40 volume percent (vol%) ethanol, but at higher HOV levelsRON was significantly higher relative to values measuredusing a modified test method with heating of the intake airto remove the HOV effect. The authors suggest that at ethanollevels above about 30%, HOV cooling is so large that the intakemixture becomes saturated and liquid droplets are beinginducted into the CFR engine. This study has been interpretedby others (but not by the original study authors) as indicatingthat the RON test captures the increased HOV from ethanolblends, and is supported by DI engine studies of fuels havingdiffering ethanol content but matched RON where no additional knock resistance was observed [13, 29, 30]. Yet thesestudies conflict with other studies cited above, and with thewell-established effect of evaporative cooling in DI enginesand the greater degree of cooling observed for ethanol blends.Sluder and coworkers [25] addressed this apparent contradiction by pointing out that in studies showing an HOVbenefit, HOV is covariant with S [22,23,26]. However, studiesthat fixed S while varying HOV did not show additional knockbenefit for higher HOV fuels [13,25,31]. Based on these studies,Sluder and coworkers suggest that HOV be considered as athermal component to S. However, S has a different effect onknock resistance at different engine operating conditions andthis may also be true for HOV. The work described here is partof a larger effort to characterize the HOV of gasoline blends,examine the effect of HOV on knock and emissions at variousengine conditions, and potentially learn how to leverageHOV to improve engine performance using specificoperation strategies.MethodsResearch Fuel PropertiesTable 1 details key properties of the fuels used in this study.Renewable oxygenates were blended into surrogates forgasoline blend-stock for oxygenate blending (BOBs) based ontoluene, isooctane, and n-heptane. Surrogate BOBs werechosen to facilitate the design of fuel blends to target 100 RON,approximately 11 S, and a range of HOVs, using informationfrom blending studies by Foong et al. [32] and Morgan et al.[33]. The two-component oxygenate blend E20 2% p-cresolwas chosen to build on results of previous studies of drop-ingasoline fuels by Ratcliff et al. [34] and McCormick et al. [35],using blending RON values developed in those studies todesign the blend. Isooctane was used as a benchmark fuel,while toluene standardization fuel (TSF) having a nominalRON of 99.8 (TSF99.8)-the terminology and compositioncome from ASTM D2699 and D2700-provided a strictlyhydrocarbon fuel with high S and low HOV. The compositionsof TSF and the TRF blends used for blending with oxygenatesare shown in Table 2. The Fuels for Advanced CombustionEngines (FACE) gasoline B is a highly paraffinic, narrowdistillation range research gasoline developed by theCoordinating Research Council [36]. It was blended with25 vol% ethanol (E25-FACE B) to provide a higher-octanenumber benchmark for the study. RON and MON wereTABLE 1 Properties of the fuel blends (HOV at 25 C).FuelRONS0HOVOxygenate Oxygen(kJ/kg) (vol%)(wt%)Isooctane10030300TSF99.8100.4 11.339000E20 2%p-Cresol-TRF88(E20)101.19.447219.7 1.57.4E40-TRF69 (E40)99.212.2 59539.414.0E25-FACE B105.6 11.8 48523.79.16 2018 SAE International; Ford Motor Company; National Renewable Energy Laboratory. SAE International2

Effects of Heat of Vaporization and Octane Sensitivity SAE InternationalTABLE 2 Surrogate BOB, standardization fuel and gasolinecompositions from detailed hydrocarbon analysis by ASTMD6729, before oxygenate %)TSF99.8 (nominally74 vol% toluene inPRF38.5)73.610.316.1TRF88 (nominally 30 vol%toluene in PRF70)28.350.421.2TRF69 (nominally 40 vol%toluene in PRF32)41.218.540.3FACE B5.9 a86.3 b7.5 cSum of aromatics, C8; b Sum of iso-paraffins; c Sum of n-paraffinsPRF: primary reference fuel.ameasured by ASTM D2699 and D2700, respectively. Based ondetailed gas chromatographic analysis using ASTM D6729,the total HOV was calculated as detailed in reference [24].Single-Cylinder EngineExperimentsSpecifications for the single-cylinder engine are listed inTable 3. This research engine was developed from a 2009model year General Motors Ecotec 2.0 L LNF-series engine,with a side-mounted DI combustion system. The dynamometer, engine control, air handling, fuel supply, combustionanalysis, and emission measurement systems have been previously described [34, 37]. The single-cylinder engine was additionally modified with an upstream fuel injector (UI) usingthe same injector type as installed in the cylinder head. Thisinjector was mounted 0.66 m upstream of the intake port andheat tape was wrapped on the intake runner wall. A thermocouple was mounted to the external surface of the intake walland this temperature was controlled to 120 C to assist fuelevaporation in UI mode. This allowed the engine to beoperated with pre-vaporized fuel, thereby negating the chargecooling effects from in-cylinder fuel evaporation that takeplace in DI mode. Intake air temperature (IAT) was controlledbased on a temperature measurement at the intake port,downstream of the UI and heated intake runner. Whenchanging fuels, a standard procedure was used for purgingthe fuel system, followed by running the engine at 2500 rpmand 9 bar net mean effective pressure (NMEP) for 30 minutes.All experiments were performed at 1500 rpm using a startof injection at 280 after top dead center (aTDC), intake valve SAE InternationalTABLE 3 Single-cylinder engine specifications.Displacement (L)0.5Bore (mm)86.0Stroke (mm)86.0Connecting rod length (mm)145.5Piston pin offset (mm)0.8Compression ratio9.2Number of valves4Combustion SystemSide-mounted DI 2018 SAE International; Ford Motor Company; National Renewable Energy Laboratory.3closing at 134 aTDC, and λ 1.0 1.01. A valve overlap of24 , beginning at intake valve opening ( 355 aTDC), waschosen to reduce trapped residuals. Note that residual gasfraction should be near zero at all knock-limited conditionsstudied, because intake manifold pressure was always higherthan exhaust pressure during the valve overlap period, so therewas no driving force to trap residual exhaust. Start of injectionat 280 was previously shown to minimize PM emissionsfrom this engine at 1500 rpm, and by inference to avoid orminimize fuel impingement on the piston or cylinder wall[37]. Air flow was measured using a critical flow orifice systemand fuel f low using a mass f low meter, as previouslydescribed [37].Two types of engine experiments were performed. Thefirst were load sweeps from 7 to 19 bar NMEP, beginning withmaximum brake torque (MBT) spark timing, then retardingthe combustion phasing as necessary to control knock. Loadsweeps were performed at a fixed IAT of 50 C. The otherexperiments measured knock-limited loads as a function ofIAT spanning 35 C to 90 C at retarded combustion phasing-afixed crank angle for 50% mass fraction burned (CA50) of20.5 aTDC. These experiments were performed using bothDI and UI modes of fueling to determine the effects of evaporative cooling from a given fuel’s HOV on knock-limitedperformance. In the UI mode, the fuel was pre-vaporizedbefore entering the combustion chamber, thus neutralizingthe charge cooling that occurs when fuel is injected directlyinto the cylinder. Knock limits were based on a knock-integral(KI) calculation, using a value of 10 for the load sweeps. Alower KI of 7.8 was used for the IAT effects study to stay withinthe engine’s maximum cylinder pressure limit of 10,000 kPa.KI was calculated by the National Instruments (formerlyDrivven) Combustion Analysis System (NI CAS) in near realtime. The method first filtered the in-cylinder pressure signalusing a second-order Butterworth band-pass filter (5-15 kHzband), then rectified the filtered signal [38]. The knock windowselected started at 5 aTDC and lasted for 30 . The referencewindow started at 210 aTDC with 30 duration. The filtered,rectified signal was then integrated for each of the knock andreference windows. The KI value was defined by:KI ò /òKnockReference- Threshold(2)The threshold value chosen for this study was 0.5 KI units.NI CAS also provides post-processing for the raw datafiles. The bulk gas temperature calculation was based on themeasured in-cylinder pressure (average of 100 cycles), slidercrank model computed volume and the temperature at intakevalve closing (IVCT). The presence of any exhaust residualfraction and the effects this would have on the specific heatratio (gamma) and IVCT were ignored. This was justifiedbased on the determination that intake and in-cylinder pressures were always higher than exhaust pressure during thevalve overlap period for knock-limited operation. For UIexperiments IVCT was assumed equal to the controlled IAT.For DI experiments IVCT was computed from the ideal gaslaw using the known cylinder volume, measured pressure,and measured air and fuel flow rates. A single-zone model,without heat transfer, was used to compute the apparent heat

4Effects of Heat of Vaporization and Octane Sensitivityrelease rates (AHRR) from the average in-cylinder pressureof 100 cycles [39,40]. Calculation of AHRR utilized a constantspecific heat ratio of 1.34 for all cases.FIGURE 2 Results of IAT sweeps using both DI and UImodes at 1500 rpm and CA50 held constant at 20.5 aTDC.Knock limit defined as KI 7.8. Fuel HOV values shown are for25 C.ResultsFigure 1 shows results for the DI load sweeps versus the CA50metric for combustion phasing. The performance benefit fromfuels that have S 11 compared with isooctane (S 0) is clear.A load of 1500 kPa NMEP was the highest output achievablefor isooctane because nearly 17 of spark retard (relative toMBT spark timing) was required to control knock. Theresulting late combustion phasing at CA50 24 aTDC justavoided combustion instability, which manifested as coefficient of variation 2% for NMEP with more retarded CA50.In contrast, only 6 of spark retard was required to controlknock for the high S fuels at 1500 kPa NMEP. Minimizingcombustion phasing retard implies higher thermal efficiencybecause the phasing is closer to optimal (i.e., MBT). Theseresults require that K in equation (1) is negative at retardedspark timing. Note that even higher loads could be achievedfrom the high S fuels by retarding combustion phasing further.However, an upper NMEP limit of 1900 kPa was imposedbecause this load corresponded with peak cylinder pressuresduring knocking cycles at or near General Motor’s recommended limit of 10,000 kPa; this threshold triggered an automatic attenuation of fuel injection rate to protect the engine.The three S 11 fuels all produced essentially the sameperformance, within experimental error. Thus, at these conditions, knock-limited performance appears to be independentof the fuels’ HOV (ranging from 390 kJ/kg to 595 kJ/kg at25 C), suggesting that HOV effects on knock resistance areincluded in S under these operating conditions. This isFIGURE 1 DI load sweeps versus CA50 combustion phasingat IAT 50 C and 1500 rpm. MBT combustion phasing wasmaintained up to the knock-limit (KI 10), and then sparktiming was retarded to maintain this value of KI as loadwas increased. SAE InternationalEngine-Fuel Performanceconsistent with the theory of HOV being a thermal contributorto S proposed by Sluder et al. [25].Figure 2 shows NMEP across a range of IATs at latecombustion phasing (CA50 held constant at 20.5 aTDC)for both DI and UI fueling. Tables of the data presented inFigure 2, along with other parameters investigated, are shownin the appendix, Tables A-1 and A-2. Comparing results forUI versus DI mode, much higher knock-limited load is achievable with DI because of charge cooling. For isooctane theincrease in NMEP ranges from 100 to 200 kPa. For TSF99.8and E20 the average NMEP difference is about 200 kPa at50 C, but increases to 250 to 300 kPa at higher temperatures.For E40 the average NMEP increase for DI is over 400 kPa at75 and 90 C IAT. Comparing the high-S fuels with isooctane,much higher loads are obtained at all IAT values, indicatingthe K in equation (1) is less than zero for all experimentsreported in Figure 2. Finally, varying HOV had little effect onload in UI mode as the high-S fuels achieved approximatelythe same load at a given IAT.The most significant observation from Figure 2 is themuch higher NMEP achieved by E40 at 75 C ( 125 kPa) and90 C ( 150 kPa) relative to E20 and TSF99.8. The E40 exhibited slightly higher S than the other fuels, however K isexpected to become more positive as IAT increases suggestingthat this effect should diminish. Significantly higher evaporative cooling of the intake charge would be expected for E40,but appears to have little effect at 50 C or lower IAT, consistentwith the proposal that HOV is incorporated into S. SAE InternationalCombustion AnalysisCylinder pressure versus crank angle results for all experiments at 90 C IAT are shown in Figure 3. Results from othertemperatures are shown in the appendix, Figures A1 - A3.Because CA50 is held constant, these plots are all very similarfor a given IAT with almost identical crank angle for peakpressure, but show a decrease in pressure with reduced load.Apparent heat release rates versus crank angle are are shown 2018 SAE International; Ford Motor Company; National Renewable Energy Laboratory.

Effects of Heat of Vaporization and Octane SensitivityFIGURE 4 a) apparent heat release rate for all fuels atIAT 90 C and b) expanded view of pre-spark region. Spark atapproximately 1 aTDC.in Figure 4 for experiments at 90 C IAT, and in the appendix,Figures A4 - A6 for other conditions. Crank angle for peakheat release and combustion duration are very similar in alltests for these fixed CA50 conditions. As previously reportedby Szybist and Splitter who used the same engine model butoperated at 2000 rpm and 20 bar IMEP [41], we observe someevidence for pre-spark heat release for isooctane in UI modeat every IAT (see appendix) and for TSF99.8 in UI mode atIAT of 90 C (Figure 4b). This observation is not the focus ofthe current paper and is not discussed further.The mixture pressure-temperature (P-T) trajectory isclosely related to K and thus to OI in equation (1). Bulk gastemperatures are a function of IVCT. Figure 5 shows pre-spark(or end of compression) temperatures for the fuels tested at90 C IAT. In UI mode differences in gamma and air-to-fuelratio lead to significantly lower end of compression temperatures for the ethanol blends than for TSF99.8. The steeplyincreasing temperature curve for TSF99.8 at end of compression is consistent with the idea that this fuel is beginning toreact pre-spark. For DI mode end of compression temperatures are significantly lower. There is only a small differencebetween E20 and E40, however this is consistent with thedifference in IVCT of 3 K for these fuels under these conditions.The fuel HOV measurements and combustion stoichiometry can be used to estimate a theoretical maximum possiblelevel of evaporative cooling [22]. The calculation must alsoreflect that HOV declines with increasing temperature asshown in Figure 6a. The calculation assumes that a stoichiometric quantity of fuel and air at IAT evaporates instantaneously with no heat transfer, and no wall wetting. Underthese assumptions the adiabatic evaporative cooling, ΔTad isgiven by equation (3):DTad m fuel * HOV / ( ( m fuel mair ) * C p-mixture )(3)mfuel is the mass of fuel, HOV is the heat of vaporization,mair is the mass of air, and Cp-mixture is the heat capacity at 2018 SAE International; Ford Motor Company; National Renewable Energy Laboratory. SAE International SAE InternationalFIGURE 3 Engine pressure data for 90 C IAT condition,CA50 fixed at 20.5 .5constant pressure of the fuel-air mixture. Calculated ΔTadresults are presented in Figure 6b. While the fuels used hereare surrogates containing only a few components rather thanfull boiling range gasolines, the calculated ΔTad values areconsistent with full boiling range gasoline results presentedin previous publications [22,24].Kasseris and Heywood, using the multi-cylinder versionof the same GM engine model used here, estimated the evaporative cooling for DI compared to PFI by determining howmuch IAT needed to be increased in DI mode to achieve thesame maximum pressure at knock onset as in PFI mode[22, 23]. The temperature increase required ranged from 70to 80% of ΔTad and was generally close to 70%. Figure 7compares IVCT from the ideal gas law based calculation with100% of ΔTad for the DI experiments. For points that fall abovethe diagonal parity line, the IVCT from ideal gas calculationis colder than that from ΔTad, while points below the diagonalare for ICVT from ideal gas that are hotter than those based

Effects of Heat of Vaporization and Octane SensitivityFIGURE 7 Comparison of intake valve closing temperaturesfor DI experiments calculated using the ideal gas law and thoseestimated from IAT and ΔTad. SAE InternationalFIGURE 5 Simulated bulk gas temperature for compressionat 90 C IAT. Spark at approximately 1 aTDC. SAE International6likely points to poor accuracy in measuring the fuel flow ratesused in the calculation. The steep slope of the plotted data isconsistent with the hypothesis that a higher fraction of ΔTadis attained a higher IAT because of more rapid evaporation ofthe fuel leading to less opportunity for heat transfer to enginesurfaces [22,23].FIGURE 6 HOV as a function of temperature for the fuelstested in this study (a), and adiabatic evaporative cooling forstoichiometric fuel-air mixtures of these fuels (b). SAE InternationalEstimation of K and OIon ΔTad. The lowest temperatures in Figure 7 are for the lowestIAT values. At low IAT (50 C and lower) the ideal gas lawcalculation predicts less evaporative cooling than ΔTad. AtIAT of 75 C the ideal gas calculation predicts cooling ofapproximately 100% of ΔTad. For IAT of 90 C, Figure 7 indicates that IVCT from the ideal gas law exhibits a slightlyhigher level of cooling than available for the adiabatic case.We have high confidence in our air flow measurement, so thisFigure 8 shows the pressure-temperature (P-T) trajectories forthe high-S fuels examined in this study at 90 C IAT, as wellas the trajectory for the RON test. The RON trajectory wastaken from [41] and is for 100 RON. These trajectories extendup to spark timing. As Figure 8 reveals, conditions for theseexperiments are at “beyond RON” conditions even for thehighest IAT case, and therefore values of K in equation (1) areexpected to be negative - consistent with the observable effectof higher S to increase fuel knock resistance.The K factor in equation (1) depends on the pressuretemperature history of the unburned gas, it is not a fuelproperty [17]. K can be estimated from experimental datausing a multiple linear regression approach that ideallyrequires a large set of fuels where RON and S are not correlated[42]. Three of the four fuels tested here have nominally thesame RON and S, violating the assumption for regression thatindependent variables be uncorrelated. Additionally, as weshow below, K is affected by evaporative cooling, leaving onlythe UI experiments as suitable for the multiple linear regression analysis approach. With only three data points availableat each IAT, the data are not adequate for attempting thisanalysis. Several other approaches were examined, includingthe PRF method via interpolation [43], but were also notworkable using this dataset.K can also be approximately predicted from the temperature when the pressure reaches 15 bar (1500 kPa) during the 2018 SAE International; Ford Motor Company; National Renewable Energy Laboratory.

Effects of Heat of Vaporization and Octane SensitivityFIGURE 9 Regression analyses for DI and UI cases of NMEPvs. OI for experiments conducted at different IAT, withCA50 20.5 aTDC. OI based on equation 4 (K4) or equation5 (K5). SAE International SAE InternationalFIGURE 8 Pressure-temperature trajectories forcompression compared to the RON test at IAT 90 C.7compression stroke (Tcomp15) as suggested by Kalghatgi [44].Two equations were proposed based on lean mixture homogeneous charge compression ignition experiments - equations(4) and (5), valid for Tcomp15 below 825 K:K 0.0056 ( Tcomp15 ) - 4.68(4)K 0.00497 ( Tcomp15 ) - 0.135 l - 3.67(5)Tcomp15 has been shown to be correlated with measured Kvalues for knocking engines as well [45, 46, 47] although it isnot applicable under all conditions [41]. However, a recentstudy using the same engine model as employed here foundthat equation (4) predicted physically reasonable values of Kthat led to OI values that were well correlated with knocklimited CA50 [41]. These results were obtained at 2000 rpm,but at similar values of IAT and at loads both above and belowthose of the current study.Tcomp15 was calculated using IVCT based on the ideal gaslaw for DI cases, while UI cases assumed IVCT IAT. Valuesof K were calculated using both equation (4) and equation (5),and all calculated K and OI values are shown in Tables A-1and A-2. Estimated K ranged from 0.7 to 1.7, in line withexpectations. In a few cases the ideal gas calculation predictsIVCT lower than the adiabatic case. The worst case is 11 Kmore cooling than predicted from ΔTad, which translates toat most a 10% difference in K from equation 5, and given themagnitude of S relative to RON only a few percent differencein OI. K calculated from equation (5) were about 0.5 unitsmore positive than K from equation (4), and produced bettercorrelations between OI and knock-limited NMEP as shownin Figure 9. 2018 SAE International; Ford Motor Company; National Renewable Energy Laboratory.While the analysis leading to Figure 9 is based on estimated rather than measured values of K, the fact that knocklimited NMEP is reasonably well correlated with OI (usingequation 5) supports the theory that HOV is a thermal component of S and thus captured in the RON test for blends up toapproximately E40 (or to an HOV equivalent to that of E40)[25,28]. The reason higher loads were achieved using E40 at75 C and 90 C IAT

2 EFFECTS OF HEAT OF VAPORIzATION ANd OCTANE SENSITIVITy 2018 A International ord Motor Company ational Renewable nerg aboratory. is, with less use of spark timing retard and fuel enrichment-i

Related Documents:

vap heat of vaporization at some standard temperature De i additive atomic contributions for the energy of vaporization Dv i additive group contributions for the energy of vaporization n number of main chain skeletal atoms X degree of crystallization V c molar volume crystalline pha

reported. A brief review of temperature adjustments of vaporization enthalpies from temperature of measurement to the standard reference temperature, 298.15 K, is included as are recently suggested reference materials. Vaporization enthalpies are included for organic, organo-metallic, and a few inorganic compounds. This compendium is the third

The heat of fusion, L f, is the heat per unit mass that is transferred in a solid-liquid phase change. The heat of vaporization, L v, is the heat per unit mass transferred in a liquid-gas phase change. The heat transf

2.12 Two-shells pass and two-tubes pass heat exchanger 14 2.13 Spiral tube heat exchanger 15 2.14 Compact heat exchanger (unmixed) 16 2.15 Compact heat exchanger (mixed) 16 2.16 Flat plate heat exchanger 17 2.17 Hairpin heat exchanger 18 2.18 Heat transfer of double pipe heat exchanger 19 3.1 Project Flow 25 3.2 Double pipe heat exchanger .

pend on the surface fluxes of heat H and moisture LE, defined in the bulk transfer formulas, H 5rc p C H U(T s T gDz) and (1a) LE5r d L y C H U(q s q). (1b) In the above equations, r d and r are, respectively, the density of dry and moist air; c p is the specific heat for air at constant pressure; L y is the latent heat of vaporization;

heat, a heat pump can supply heat to a house even on cold winter days. In fact, air at -18 C contains about 85 percent of the heat it contained at 21 C. An air-source heat pump absorbs heat from the outdoor air in winter and rejects heat into outdoor air in summer. It is the most common type of heat pump found in Canadian homes at this time.

HEAT CHANGE OVER VALVE B C R L2 L1 (HOT) Typical 3H/2C or 2H/1C Heat Pump System System: Indicates current mode of operation. AUXILIARY HEAT RELAY EMERGENCY HEAT RELAY Terminal 2 Heat 2 Cool Conventional System 2 Heat 2 Cool Heat Pump System 3 Heat 2 Cool Heat Pump System RC RH C B O G W/E W2 Transformer power (cooling) Transformer power .

An Introduction to Random Field Theory Matthew Brett , Will Penny †and Stefan Kiebel MRC Cognition and Brain Sciences Unit, Cambridge UK; † Functional Imaging Laboratory, Institute of Neurology, London, UK. March 4, 2003 1 Introduction This chapter is an introduction to the multiple comparison problem in func-