You are on page 1of 22

International Journal of Engine Research http://jer.sagepub.

com/

Characterizing the thermal sensitivity of a gasoline homogeneous charge compression ignition engine with measurements of instantaneous wall temperature and heat flux
J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt and R Rask International Journal of Engine Research 2005 6: 289 DOI: 10.1243/146808705X30558 The online version of this article can be found at: http://jer.sagepub.com/content/6/4/289

Published by:
http://www.sagepublications.com

On behalf of:

Institution of Mechanical Engineers

Additional services and information for International Journal of Engine Research can be found at: Email Alerts: http://jer.sagepub.com/cgi/alerts Subscriptions: http://jer.sagepub.com/subscriptions Reprints: http://www.sagepub.com/journalsReprints.nav Permissions: http://www.sagepub.com/journalsPermissions.nav Citations: http://jer.sagepub.com/content/6/4/289.refs.html

>> Version of Record - Aug 1, 2005 What is This?

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

SPECIAL ISSUE PAPER

289

Characterizing the thermal sensitivity of a gasoline homogeneous charge compression ignition engine with measurements of instantaneous wall temperature and heat ux
J Chang1, Z Filipi*1, D Assanis1, T-W Kuo2, P Najt2, and R Rask2 1Mechanical Engineering, University of Michigan, Ann Arbor, Michigan, USA 2General Motors R&D, Warren, Michigan, USA The manuscript was accepted after revision for publication on 10 May 2005. DOI: 10.1243/146808705X30558

Abstract: An experimental study was performed to provide qualitative and quantitative insight into the thermal eects on a gasoline-fuelled homogeneous charge compression ignition (HCCI) engine combustion. The single-cylinder engine utilized exhaust gas rebreathing to obtain large amounts of hot residual gas needed to promote ignition. In-cylinder pressure, heat release analysis, and exhaust emission measurement were employed for combustion diagnostics. Fast response thermocouples were embedded in the piston top and cylinder head surface to measure instantaneous wall temperature and heat ux, thus providing critical information about the thermal boundary conditions and a thorough understanding of the heat transfer process. Two parameters determining thermal conditions in the cylinder, i.e. intake charge temperature and wall temperature, were considered and their eect on ignition and burning rate in an HCCI engine was investigated through systematic experimentation. The approach allowed quantitative analysis, and separating qualitatively dierent eects on the core gas temperature from the eects of near-wall temperature stratication. The results show great sensitivity to changes in wall temperature and such like, but a somewhat weaker eect of intake charge temperature on HCCI combustion. Variations of combustion phasing and peak burn rates due to wall temperature changes can be compensated if the intake charge temperature is varied in the opposite direction and with a factor of 1.11. The combustion stability limit of the HCCI engine depends more on wall temperature than on intake charge temperature. Analysis of a large number of individual cycles indicates that decreasing intake temperature retards timing, and the burn rates change primarily as a function of ignition timing. In contrast, lowering the wall temperature led to greater reduction in the bulk burn rate and greater increase in cyclic variability than expected simply as a result of retarded ignition, thus indicating signicance of the thermal stratication in the near-wall boundary layer. Keywords: gasoline HCCI, heat transfer, wall temperature, combustion, cyclic variability 1 INTRODUCTION Energy supply concerns and increasingly stringent exhaust emissions regulations require new propulsion systems that provide high performance, excellent fuel economy, as well as reduced exhaust emissions. The spark-ignition engines combined with a three-way catalyst are currently the mainstay
* Corresponding author: Mechanical Engineering, The University of Michigan, W E Lay Automotive Laboratory, Ann Arbor, MI 48109-2121, USA. email: lipi@engin.umich.edu

of the US automotive market, owing to the extremely low tailpipe emissions and favourable production costs. However, technology and cost constraints pose a challenge with respect to meeting future targets. Wider application of compression ignition (diesel) engines to light vehicles is promising, owing to their high fuel eciency. Unfortunately, locally high combustion temperature and diusion combustion characteristics lead to elevated formation of NO and x soot, and aftertreatment technology requires signicant improvement before it can approach the eciency of the three-way catalyst in the SI engine
Int. J. Engine Res. Vol. 6

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

JER03005 IMechE 2005

290

J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt, and R Rask

application. Homogeneous charge compression ignition (HCCI), also known as controlled autoignition (CAI) engine when augmented with an additional source of thermal energy, has emerged as a concept potentially marrying diesel-like eciency with reduced emissions of NO and soot. Unthrottled, x lean operation with an elevated compression ratio ensures high eciency, while spontaneous combustion of a premixed lean charge eliminates local high-temperature, rich sources of NO and soot. x Realizing the potential benets of HCCI combustion depends on resolving the main challenge, which is combustion control without the direct trigger and management of thermal transients. The timing of HCCI combustion is determined by the chemical kinetics and the autoignition can be aected only indirectly through overall control of conditions in the cylinder at the end of compression. Two factors are especially important: the mixtures chemical composition, and the in-cylinder thermal environment. In-cylinder temperature history is critical for ensuring the compression ignition of the charge and determining its timing, given the charge composition. Hence, a thorough understanding of the eect of thermal conditions on HCCI energy release rate can signicantly aid eorts in devising strategies for eective HCCI combustion control during steady operation and thermal transients. Heat transfer through the combustion chamber surface plays an important role in determining the thermal conditions of mixture. Subsequently, surrounding wall temperatures in the cylinder could change the ignition and burning rate considerably by changing the thermal condition of the charge adjacent to the wall. While the signicance of gas thermal state and heat transfer in the HCCI engine is obvious, there is little available data to quantify this eect or shed light on fundamental questions, such as the role of the core gas temperature versus the temperature gradients and thermal boundary layers near the wall. Combustion chamber surface temperature measurement in an HCCI engine, combined with combustion diagnostics, can provide invaluable data for understanding the above phenomena and subsequently for deriving proper management of hot-to-cold and cold-to-hot transients. In particular, this paper is focused on determining the sensitivity of main in-cylinder processes (e.g. ignition, combustion, emission formation, heat ux) to changes of thermal conditions of the charge or the wall through systematic experimentation. Surface heat ux sensors are installed on the cylinder head and the piston to provide the local wall temperature and heat ux information under varying coolant and
Int. J. Engine Res. Vol. 6

oil temperatures, the intake charge temperature or both. As the coolant temperature is reduced, the HCCI engine begins to encounter unstable combustion, and an additional goal of this work is to determine the increase of the intake temperature required for compensating the eect of cooler walls. Detailed local instantaneous temperature and heat ux measurements are performed to provide insight into heat transfer characteristics and, particularly, thermal stratication in the combustion chamber. Several previous studies considered the eect of the intake charge temperature on the HCCI burn rate [16]. They generally concluded that a higher intake charge temperature advances ignition timing and shortens burn duration, eventually enabling an extension of the lean airfuel ratio limit [13, 5]. Higher charge temperature speeds up the rate of pre-ignition reactions and enables the conditions critical for ignition to be reached sooner. However, the sensitivities of combustion to intake charge temperature were dierent depending on the engine size, compression ratio (CR), fuel type, and HCCI enabling strategies (e.g. negative valve overlap or high CR). For instance, Dec et al. [7] tested a high compression ratio engine and showed a dierent intake temperature requirement for dierent types of fuel, such as isooctane, gasoline, and primary reference fuel. The results from Persson et al. [4] indicate lower impact of the intake charge temperature in an engine using a high amount of hot internal residual than in a high CR engine [7], and much higher sensitivity to residual temperature. Higher coolant temperature extends the lean operating limit [1]. In contrast, Sjoberg et al. [6] suggest lowering the intake charge temperature as a means of smoothing heat release rate and broadening the high load range of HCCI. Whereas Sandia researchers [6] provide very valuable insight into phenomena occurring in the high compression ratio lean HCCI engine with conventional gas exchange process, there are no previous studies focused on fully characterizing thermal eects on the engine with rebreathing, including comparison between the eects of intake charge and coolant temperature. In addition, there are no published direct measurements of instantaneous surface temperature or heat ux to support the analysis and characterization of thermal eects on HCCI combustion; hence the specic motivation for the approach taken in this work. The engine used in this study is a single-cylinder engine operated with a premixed charge of air and gasoline. The engine operates with rebreathing, i.e. the exhaust valve is opened one more time during the intake stroke to allow induction of hot residuals
JER03005 IMechE 2005

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

Characterizing the thermal sensitivity of a gasoline HCCI engine

291

and to help promote autoignition. The coaxial-type fast-response thermocouples (TC) are installed ush with the surface of the piston and the cylinder head for measuring instantaneous temperature as a function of crank angle. Processing of instantaneous temperature signals using unsteady heat conduction equation and Fourier analysis provide local heat ux proles. The paper is organized as follows. The experimental engine set-up is described in section 2. Methodology for local surface temperature and heat ux measurements including error analysis is discussed in the same section. The eects of independent variations of intake charge temperature and coolant temperature on combustion, performance, and emissions are investigated next, and the thermal phenomena responsible for observed dierences between the two are analysed using local surface temperature and heat ux measurements. A similar type of combustion and heat ux analysis is then applied to a study varying the intake and coolant (wall) temperature dependently, such that combustion phasing remains constant. In this manner, the variations of heat ux as a result of near-wall thermal stratication are separated from the eect of burn rates as the main driver of the transient heat transfer. The last section oers a summary and conclusions.

2 EXPERIMENTAL ENGINE SET-UP A modied single-cylinder engine is used for engine dynamometer tests. A GM prototype, pent-roofshape cylinder head is designed especially for this experiment, but basic features correspond to a typical modern four-valve cylinder head. Table 1 provides engine specications. An exhaust rebreathing strategy is applied to provide the necessary internal residual fraction for HCCI combustion control. This involves the exhaust valves opening for a second time during the intake stroke via an additional lobe on the

Table 1 Engine specications


Engine type Bore/stroke Displacement Connecting rod length Compression ratio IVO/IVC Main EVO/EVC 2nd EVO/EVC Fuel type 4 valves, single cylinder 86.0/94.6 mm 0.549 l 152.2 mm 12.5 346/592 * 130/368 * 394/531 * Gasoline (H/C=1.898)

*0 crank angle is assigned to TDC combustion. Intake valve open/intake valve close. Exhaust valve open/exhaust valve close.

exhaust cam. During this period, hot exhaust residual gas is drawn back into the cylinder subsequently to help with ignition during HCCI operation. The hot residual aects autoignition and combustion through a thermal and chemical eect. The exhaust back pressure is maintained at the level close to ambient pressure by controlling the gate valve shown in Fig. 1. One of the intake ports has a tangential orientation and is intended to produce an organized swirling motion, while the other contains a swirl control valve (SCV) installed ~180 mm upstream from the valve seat (see Fig. 1). The SCV was originally devised to control the swirl/tumble intensity in the engine. However, the addition of the second lobe on the exhaust cam and the use of the exhaust rebreathing strategy makes the ow pattern in the HCCI engine much less organized and reduces the impact of the SCV. Exhaust gas is sampled at the exhaust plenum and gaseous analysers measure concentrations of HC, NO , CO, CO , and O , thus x 2 2 allowing calculation of engine-out airfuel ratio and combustion eciency. The high-pressure fuel delivery system is based on a bladder-type accumulator. A bladder containing fuel is in a vessel pressurized by nitrogen. Regulating the pressure of N allows control of the fuel injection 2 pressure. The fuel delivery line directs the fuel to a direct-injection-type injector in the vaporizer for fully premixed operation. In this work, all tests were performed with a fully premixed charge in order to ensure compositional homogeneity of the mixture. The fuel vaporizer consists of a 320 cm3 aluminium chamber surrounded by a 750 W electric band heater to maintain the inner surface temperature at ~220 C, sucient to fully boil gasoline at 100 kPa. The vaporizer air intake is positioned above the main throttle valve (see Fig. 1), similar to the design proposed by Dec et al. [7], thus ensuring the pressure drop required for the desired air ow. The air ow is great enough to prevent condensation in cold spots and stable operation, and yet small enough to keep the mixture well outside of the ammable limit. Residence time of fuel in the vaporizer has been estimated at ve engine cycles. Vaporized fuel and air are routed to the intake runner upstream of the plenum to ensure complete mixing. Mixture temperature is controlled by the electric heater located upstream of the intake plenum. A Kistler 6125A piezoelectric pressure transducer measures the pressure trace in the cylinder with 0.5 crank-angle resolution. A ame arrester is installed in front of the transducer tip to avoid thermal shock, and 200 consecutive cycles are recorded at any given condition. Pressure signal ltering is performed,
Int. J. Engine Res. Vol. 6

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

JER03005 IMechE 2005

292

J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt, and R Rask

Fig. 1 Schematic of HCCI engine set-up

based on a SavitzkyGolay ltering scheme [8]. To prevent phase shift after ltering, the ltered sequence is then reversed and run back through the lter. Top dead centre is determined by considering the thermodynamic loss angle via the net heat release method suggested by Sta [9]. Absolute cylinder pressure referencing, or pressure pegging, is made using intake manifold absolute pressure [10]. The heat release analysis is performed, based on in-cylinder pressure measurements. An in-house developed single-zone heat release model is used for the analysis. Detailed model explanations are given by Chang et al. [11]. The results of heat release analysis provide the crankangle resolved net energy release rate, as well as the cumulative energy release.

3 LOCAL WALL TEMPERATURE AND HEAT FLUX MEASUREMENT 3.1 Instantaneous temperature measurement methodology A variety of surface thermocouple designs has been developed since Eichelbergs rst attempt to measure instantaneous surface temperature in the combustion chamber in 1939 [12]. A surface junction on modern probes can be formed by very thin metallic layers vacuum deposited or chemically plated to obtain fast response time. Fast response thermocouples can be categorized in the following main
Int. J. Engine Res. Vol. 6

groups: the coaxial, the pair wire, and the thin lm type. The former two are more widely used for measurements on metal surfaces, while the thin lm design has been developed specically for measurements on the surface of ceramic coatings and inlays [13, 14] with the intention of minimizing the disturbance in the temperature eld owing to the insertion of the probe. A coaxial-type probe was selected for this study owing to its evident favourable dynamic behaviour when applied to heat ux measurements during the compression and ring conditions [15, 16] in a metal engine. It consists of a thin wire of one TC material (constantan) coated with a ceramic insulation of high dielectric strength, swaged securely in a tube of a second TC material (iron). A vacuumdeposited metallic plate forms a metallurgical bond with the two TC elements, thus forming the TC junction with 12 mm thickness over the sensing end of the probe. The probes are custom manufactured for engine experimentation and their response time is of the order of a microsecond [3]. The sensing area at the tip is mounted ush with the combustion chamber surface. The main contribution to the total heat ux during ring conditions is the transient component of heat ux, and hence the accuracy of the surface thermocouple is critical. As long as the size of the hot junction is small enough, in this case its diameter is only 1.52 mm, the temperature dierence between the ends of the centre wire and the outer tube may
JER03005 IMechE 2005

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

Characterizing the thermal sensitivity of a gasoline HCCI engine

293

be neglected [15]. The penetration of the thermal pulse into the surface is small in comparison with the relevant surrounding dimensions (at/L2%1) and a one-dimensional transient heat ux analysis can be used [8, 1517]. Theoretical analysis based on the nite element method [16], and experiments in the dynamic test rig [15] indicate a high accuracy of instantaneous temperature measurements (~98 per cent), good ability of the probe to capture an instantaneous rise of heat ux, but also the potential for overpredicting the peak heat ux owing to dynamic phenomena within the probe, i.e. the inability of the heat sink surrounding the centre wire to cool it fast enough to control the transient response. The error is minimized if correct thermal properties of the probe body are used; in this case conductivity is k=57.11 (W/mK) and diusivity a=16.26E6 (m2/s). Another potential source of error is the presence of combustion deposits on the surface of the probe, and this can oset the tendency to overpredict the peak heat ux in a practical engine application. In addition, the peak heat ux is not the main quantitative indicator of the heat transfer phenomena; rather, it is the combination of overall features of the crankangle resolved heat ux prole and the integrated heat loss that has critical impact on the engine cycle. Hence, the generally accepted practice of using a coaxial-type fast-response thermocouple for measurements on the metal surface is here relied upon, but overall accuracy of measurements is also checked and veried through a heat release analysis described in [11]. Finally, a second thermocouple, or backside junction is located 4 mm below the

surface to measure in-depth temperature and allow evaluation of the steady state heat ux. It is much more dicult to ensure one-dimensional heat ow towards the backside junction than at the surface. However, given the fact that the steady component is much smaller than the unsteady term during the combustion event, this does not have a great eect on overall accuracy. As shown in Fig. 2(a), a total of nine heat ux probes, including two probes in the cylinder head and seven probes in the piston (four inside the piston bowl and three at the periphery), are installed to measure local instantaneous temperature and heat ux variations. Measurement locations on the cylinder head are dictated by available space for installing special sleeves for mounting fast-response thermocouple probes. However, locations on the piston top are strategically chosen to provide complete insight into spatial variation of instantaneous heat ux owing to local combustion dierences or possible thermal stratication. The thermocouples on the piston are embedded directly into the piston crown material. Headside TC signals are routed directly outside the engine and joined with a reference (cold) junction at ambient temperature. In contrast, TC wires from the piston surface are routed to an isothermal plate installed on the inner surface of the piston skirt. In this case, a new reference junction is located at the isothermal plate and its temperature is measured by a thermistor. This allows the use of highly durable stainless steel braided wire for transferring signals through the telemetry linkage and into the data acquisition system.

Fig. 2 Heat ux measurement instrumentation: (a) locations of heat ux probes; (b) mechanical linkage telemetry system
Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

JER03005 IMechE 2005

Int. J. Engine Res. Vol. 6

294

J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt, and R Rask

A mechanical telemetry system is used for conveying signals from the moving piston and connecting rod. Figure 2(b) shows a picture of the in-house designed grasshopper linkage, made of aluminum, prior to installation. Its geometry was optimized via a kinematics simulation [18]. Joints are designed for high lateral stability and the oating heat-treated pins are used for increased strength and reduced wear. Special electrical connectors with a total of 28 pins are custom built for the big end of the connecting rod, to allow easy installation and disassembly while ensuring reliable transmission of signals. 3.2 Instantaneous heat ux calculation The total instantaneous surface heat ux is calculated by solving the unsteady heat conduction equation with two temperature boundary conditions and one initial condition. Heat transfer is assumed to be strictly one-dimensional, and normal to the wall surface. Opris et al. [19] conducted computer simulations of piston thermal loads and concluded that heat ow through the piston is highly onedimensional, except on the sharp edges. The solution of the unsteady heat conduction equation can be obtained by applying Fourier analysis, electrical analogy, or numerical nite dierence methods, as documented in reference [15]. For this work, Fourier analysis method is chosen. The solution provides a time-dependent temperature prole at the surface which is described as
T(t)=T + S[A cos(nvt)+B sin(nvt)] (1) m n n where T is the time-averaged surface temperature, m v is angular velocity, A and B are Fourier coeffin n cients, and n is the harmonic number. A fast Fourier

transform (FFT) is applied to the measured surface temperature data to determine coecients A and n B . When performing an FFT to determine A and B , n n n the selection of the value of the harmonic number n aects results signicantly. In particular, if the number is too small, the periodic temperature prole from equation (1) will not match the measured wall temperature accurately. If n is set to be too large, calculation sensitivity to the raw data noise increases and could lead to non-physical uctuations in the simulated prole. Overbye et al. [20] used n=72 with 5 crank-angle resolution, but the authors found that n=40 is appropriate for higher resolution (0.5 crank angle) cases. Once A and B are determined, n n Fouriers law is applied to equation (1) to obtain total heat ux. The total heat ux consists of a time-independent steady term and a time-dependent transient term, which is the uctuation of heat ux at the surface, i.e. k N q = (T T )+k w [(A +B ) cos(nvt) n n n l u l m n=1 (A B ) sin(nvt)] (2) n n where w =(nv/2a)1/2. n Calculation of local heat ux at each position is performed, based on instantaneous surface temperature (T ) and backside temperature (T ) measurew l ments. An example of instantaneous temperature measurements obtained at the surface and at the backside, 4 mm below, during a sequence of 200 consecutive cycles is shown in Fig. 3(a). Heat ux proles obtained from the temperature measurements are given in Fig. 3(b). For evaluation of main heat ux features, a single prole generated as an ensemble average from 200 cycles is required for each location. However, this processing methodology presents a

Fig. 3 Three-dimensional plots of (a) instantaneous surface-side and backside temperature; (b) calculated heat ux during 200 consecutive cyclescylinder head location
Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

Int. J. Engine Res. Vol. 6

JER03005 IMechE 2005

Characterizing the thermal sensitivity of a gasoline HCCI engine

295

very heavy computational burden, given a total of nine locations. To reduce the computational eort, the ensemble average instantaneous temperature prole is found rst, and heat ux is then calculated from the single mean temperature prole. Figure 4 compares two heat ux traces at the same cylinder head location: one is determined as an ensemble average of 200 individual heat uxes, and the other is derived from an averaged temperature signal. The two proles are almost identical, thus conrming validity of the second approach. Limits in Fig. 4 indicate the bounds of a crank angle resolved standard deviation of heat ux for 200 cycles. Standard deviation of the heat ux is the highest at the peak heat ux point as a consequence of cycle-to-cycle variations of combustion. 3.3 Assessing the accuracy of instantaneous surface temperature and heat ux measurements Measurement accuracy of heat ux probes depends on their time constant, manufacturing tolerance, bias error, and precision error. All of these factors need to be considered before the measurements can be used with condence for physical analysis. Heat ux probes time constant aects the temperature signal gradient for given excitation, and hence the shapes of the heat ux prole. The time constant is dened as the theoretical time to reach 63.2 per cent of a step change in surface temperature. The nominal value of the tested vapour deposited, two-micron thick junction provided by the manufacturer is one microsecond. This is considered fast enough for crank angle resolved measurements with half-adegree resolution at up to 3000 r/min. The measure-

ment bias error aects the absolute value of surface temperature rather than the gradient of temperature uctuations or heat ux. Bias could stem from the TC itself or from the reference temperature reading. The J-type thermocouple (TC) used here satises ISA standard limits of error of 0.75 per cent. A negative temperature coecient (NTC) glass-encapsulated bead thermistor is used to monitor the reference temperature of the isothermal plate on the piston, and its accuracy is between 0.001 C and 0.01 C. Consequently, the total surface temperature bias error is insignicant, particularly in the context of the heat ux analysis. The precision error of the probe is considered to be a measure of the random variation encountered during repeatability trials. A series of motoring tests with individual cycle temperature measurements over 200 cycles provided in situ evaluation of precision, based on the assumption that the precision error can be estimated to be less than the standard deviation of temperature during motoring, since the observed variations include possible physical eects as well. The results indicate that the precision error of the probe is less than 0.37 C [21]. Hence, it can be concluded that the thermocouples provide a sucient level of accuracy for studies planned in this work, and in particular for assessments of qualitative features of heat ux proles.

4 RESULTS VARYING T AND T intake coolant INDEPENDENTLY 4.1 The eect of wall and intake charge temperature change on HCCI combustion In this section, temperature of the intake charge and the wall are varied independently, e.g. intake charge temperature is kept constant during the wall temperature variation, and vice versa. Engine tests were conducted within the acceptable wall and intake charge temperature ranges to determine the sensitivity of the combustion process to their respective changes. Combustion chamber wall temperature is varied by controlling the coolant and oil temperature. Table 2 shows operating conditions for each of the two parametric studies. Since both the intake charge temperature and wall temperature sweeps aect the in-cylinder charge temperature, volumetric eciency changes too as a result of variation of density. Hence, the decrease of airow brings a slight change of airfuel ratio from 21 : 1 to 19 : 1 when the intake charge temperature increases. No compensation of airow is made during this specic test
Int. J. Engine Res. Vol. 6

Fig. 4 Heat ux calculation procedures: ensemble average of heat uxes from 200 cycles and heat ux from ensemble average of temperatures (marks)
JER03005 IMechE 2005

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

296

J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt, and R Rask

Table 2 Operating conditions for wall temperature and intake charge temperature sweeps
Wall temperature sweep
Air fuel ratio Internal EGR Intake charge temperature Oil-in temperature Coolant-in temperature Fuelling rate Speed Swirl control valve External EGR Intake manifold pressure Exhaust manifold pressure (%) (C) (C) (C) (mg/cycle) (r/min) (deg) (%) (kPa) (kPa) 20 39 90 (xed) 8194 8095 11.0 2000 36 (partially open) 5.3 95.0 101.0

Intake charge temperature sweep


1921 3740 76116 95 (xed) 95 (xed)

because rstly, the airow control using throttle positioning is dicult when changes are subtle, and secondly, the variation of airfuel ratio is considered to be one of the realistic eects of temperature modication. The low temperature limit in each test is determined by high cyclic variation. Coecient of variation (COV) of indicated mean eective pressure (i.m.e.p.) is used as an indication of combustion stability. A total of 200 cycles of pressure data are sampled for the calculation. From the plot in Fig. 5(a), testing with less than 80 C coolant temperature or less than 75 C intake charge temperature should be avoided owing to COV of i.m.e.p. increasing over 3 per cent. In particular, COV of i.m.e.p. increases exponentially as the coolant temperature drops below 80 C. This implies that running the HCCI engine with rebreathing during the warm-up stage may be limited, and it will require proper thermal management to ensure combustion stability. The decrease of coolant temperature shows a stronger eect on the stability of combustion than the decrease of intake charge temperature. Details regarding changes of burn rates will be discussed in the second part of this section. Performance and fuel economy graphs are shown in Figs 5(b) and (c). Net mean eective pressure (n.m.e.p.) reaches a maximum over the 90105 C intake charge temperature (T ) range, and decreases beyond 105 C simply intake because combustion phasing becomes too advanced. The n.m.e.p. varies in the range of 3 per cent with respect to 40 C intake charge temperature swing. In contrast, only 15 C of the coolant temperature (T ) change produces the same n.m.e.p. variation. coolant Net specic fuel consumption (NSFC) varies almost 3 per cent under the same circumstances. Thus, thermal stratication resulting from reduced wall temperature has a stronger impact on HCCI combustion, performance, and fuel economy than intake temperature change.
Int. J. Engine Res. Vol. 6

Fig. 5 Variations of (a) COV of i.m.e.p.; (b) n.m.e.p.; (c) NSFC, with coolant temperature (square mark) and intake charge temperature (diamond mark)

Sensitivities of exhaust emissions to temperature changes are illustrated in Fig. 6. Unburned hydrocarbon (HC) emission index normalized by fuel mass shown in Fig. 6(a) increased with decreased
JER03005 IMechE 2005

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

Characterizing the thermal sensitivity of a gasoline HCCI engine

297

Fig. 6 Variations of emissions with coolant temperature (square mark) and intake charge temperature (diamond mark): (a) HC; (b) NO ; x (c) CO emissions index

temperature. HC emission increases as either the intake or coolant temperatures decrease, owing to quenching in the crevices and near the wall and consequent partial burns. HC emission almost doubles, over the range of intake charge temperature variations (11575 C), and around 45 per cent when the coolant temperature changes from 95 to 80 C. The steeper slope of the HC change versus coolant temperature can be attributed to enhanced thermal boundary layer growth and incomplete burning near the wall. Figures 6(b) and (c) show that nitrogen oxide (NO ) decreases and carbon monoxide (CO) x increases systematically as T or T decreases. intake coolant Sensitivity of the NO and CO concentration to x variations of T or T are similar, but it is intake coolant worth noting that NO concentrations are extremely x small for moderate levels of intake or coolant temJER03005 IMechE 2005

peratures. Smoke was measured by an AVL 415S variable sampling smoke meter at the exhaust plenum, and there was no detectable amount of smoke under any operating conditions. Burn rate variations under the wall temperature sweep and intake charge temperature sweep are shown in Fig. 7. Each curve is calculated from an ensemble average of 200 cycles. The results indicate that both ignition timings and burn durations are extremely sensitive to coolant temperature and intake charge temperature variations. Combustion phasing is retarded and total burn duration increases when either the wall temperature or intake charge temperature are decreased, but features of burn rate dier signicantly from one case to the other. The link between ignition timing and peak burn rates seems to be stronger in case of intake temperature variations (Fig. 7(b)), while curves in Fig. 7(a) demonstrate small variations of ignition accompanied by relatively large changes of burn rate proles. Hence, peak burn rate decreases 2.3 per cent for a 1 C coolant temperature decrement, while the intake charge temperature reduction causes a smaller variation of 1.3 per cent per C. Figure 8 shows the behaviour of the calculated mass averaged in-cylinder gas temperature near the ring TDC. In the case of coolant temperature sweep, the gas temperatures during compression (1020 BTDC) do not show signicant variation (Fig. 8(a)), since temperature dierences between the intake charge and the wall are relatively small. However, after combustion starts, slower burning in the case of lower T leads to retarded coolant phasing of the peak temperature and its reduced magnitude. The latter is due to the eects of piston motion and volume change, as well as possible temperature stratication near the wall. In the case of lowering the intake charge temperature, the gas temperature before TDC in Fig. 8(b) varies, and causes tangible autoignition timing delay and extended burn duration that seems to be coupled to phasing of ignition. Further insight into the dierent character of the intake and coolant temperature eects on combustion is provided through heat release analysis of a large number of individual cycles discussed in the next section. 4.2 Individual cycle heat release analysis correlating the ignition timing with bulk burning The results shown in the previous section suggest qualitatively dierent eects of intake charge temperature and coolant temperature on HCCI combustion, particularly regarding the correlation between
Int. J. Engine Res. Vol. 6

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

298

J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt, and R Rask

Fig. 7 Rate of net energy release behaviour with respect to varying: (a) coolant temperature; (b) intake charge temperature

Fig. 8 In-cylinder gas temperature behaviour with respect to (a) varying coolant temperature; (b) intake charge temperature

ignition timing and burn rate phasing. To investigate this issue further, the mass fraction burned curves are obtained for all 200 cycles through heat release analysis of individual cycles, using methodology described in [11]. The main combustion characteristics obtained for individual cycles are then correlated in scatter plots shown in Figs 911. The trends are evaluated with respect to the coolant temperature variation (a), and intake charge temperature variation (b). Figure 9 illustrates the relationship between the locations of 10 per cent mass fraction burned (MFB) and 50 per cent mass fraction burned point, the former being an indication of ignition timing, and the latter considered to be an indication of overall combustion phasing. Lowering either the coolant temperature or the intake charge temperature retards the ignition timing owing to the eect on chemical kinetics, but the range of change is wider in the case of T variations. Retarded intake ignition timing delays the main combustion phase
Int. J. Engine Res. Vol. 6

through eects of piston movement and volume change, but the scatter of individual points is smaller in the case of intake charge temperature variations. In particular, regression analysis produces lines having similar slopes, but the coecient of correlation (R) has higher values in the case of intake temperature change (R=0.985) than the coolant temperature change (R=0.903). Figure 10 correlates ignition timing (10 per cent burned) with burn duration (1090 per cent burn) duration. In the case of varying the intake charge temperature (Fig. 10(b)), the burn duration is strongly coupled to changes of ignition timing, and the quadratic regression curve drawn through all the points captures the main trend with the R value of 0.918. In the case of coolant temperature change (Fig. 10(a)) however, the correlation is much weaker. An attempt to nd the quadratic regression for this case produces a curve with an R value of only 0.722. In other words, the main burn duration slows down beyond
JER03005 IMechE 2005

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

Characterizing the thermal sensitivity of a gasoline HCCI engine

299

Fig. 9 Scatter plots of 10 per cent versus 50 per cent mass fraction burned data sets obtained for dierent values of (a) coolant temperature; (b) intake charge temperature

Fig. 10 Scatter plots of 10 per cent mass fraction burned versus 1090 per cent burn duration data sets obtained for dierent values of (a) coolant temperature; (b) intake charge temperature. A second order regression is determined for all points in data sets considered

Fig. 11 Scatter plots of 10 per cent mass fraction burned versus 5090 per cent burn duration data sets obtained for dierent values of (a) coolant temperature; (b) intake charge temperature. A second order regression is determined for all points in data sets considered
Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

JER03005 IMechE 2005

Int. J. Engine Res. Vol. 6

300

J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt, and R Rask

the change dictated by the retarded ignition timing as T decreases, and the clouds of points correcoolant sponding to a single T seem to be staggered and coolant shifted vertically. This is conrmed by the regression curves obtained for individual groups of points and superimposed as dashed lines on the plot in Fig. 10(a). The scatter of points increases with decreased coolant temperature, indicating larger cycle-to-cycle variations. The qualitative dierences in the way T and intake T (T ) aect the combustion event are concoolant wall rmed when 5090 per cent mass fraction burned data as a function of ignition timing (10 per cent burned) are examined in Fig. 11. Regression curve represents the overall trend very well as the intake charge temperature changes (coecient of correlation higher than 0.91), while in the case of coolant temperature variations the correlation is much weaker (R=0.75). This indicates that a decreasing coolant temperature extends late burning beyond that directly consequent on retarded ignition. Therefore, it can be concluded that an increased thermal boundary layer region owing to the colder wall temperature has a signicant eect on bulk burning, with a smaller impact on ignition through an indirect eect on core gas temperature. In contrast, the variation of intake charge temperature seems to aect all combustion characteristics through changes of core (bulk) gas temperature, leading to closely correlated variations of ignition timing and combustion duration. In summary, although the overall sensitivity of combustion and performance to coolant temperature appears to be similar in magnitude to the intake temperature eect, the thermal stratication between the core and boundary layer makes bulk burning more sensitive to variations of wall temperature, especially in an engine with a relatively small combustion chamber. 4.3 Instantaneous surface temperature and heat ux variations with coolant temperature The eect of coolant temperature change on cylinder wall surface temperature is illustrated in Fig. 12. Each plot shows seven temperature signals on the piston top and two on the head surface. Each line represents a 200-cycle ensemble averaged surface temperature swing. The hottest spot is the centre of the piston surface, and it is 63 C higher than the coolant temperature. The coldest measured spot on piston surface is the location number 5. Overall, the area with higher surface temperature is located on the intake side of the piston bowl, while the lower temperature locations are on the periphery, particularly
Int. J. Engine Res. Vol. 6

on the side with the elevated nose. The local surface temperature distribution is quite dependent on the piston geometry, particularly with regard to piston material distribution between the bowl and the squish areas. For instance, location number 6 is located at the squish area and its temperature is higher than in the vicinity. The temperature dierence between the two head locations can be attributed primarily to the coolant ow direction. The local surface temperatures decrease linearly with decreasing coolant temperature, and the spatial temperature spread from maximum to minimum point is consistently approximately 10 C for seven piston locations. Figure 13 shows four dierent plots illustrating local heat ux behaviour with respect to the coolant temperature variation. Local heat uxes obtained for T =95 C indicate that their proles are coolant relatively uniform, displaying the same phasing and very similar main features. While there is some variability in absolute magnitude, the spatial variations are much less than in case of a conventional sparkignited gasoline engine, where ame propagation causes great variation between local conditions close to the spark or at the periphery of the chamber [11, 21, 22]. This conrms the subjective impression from examining Fig. 12 that, while local temperature levels vary from location to location, the shape of the uctuations during combustion are very similar (see a plot for 95 C T in Fig. 13) when operating coolant with a compositionally homogeneous mixture. As soon as coolant temperature decreases, local heat ux decreases as well. While this might sound counterintuitive, it conrms that combustion is the main driver in the transient heat transfer process, and hence the proles follow the trend of burn rate variations with T (see Fig. 7(a)). However, peaks coolant of the proles at the periphery of the piston decrease more than at other locations an indication of the uneven growth of the thermal boundary layer with reduced coolant temperature. 4.4 Average wall temperatures on the piston and head surface Overall average wall temperature trends are shown in Fig. 14, in order to quantify the overall eects of coolant temperature and intake charge temperature on the combustion chamber wall. To represent cycleaveraged values at each specic location, surface temperatures from Fig. 12 are averaged throughout the cycle (720 crank angle) to produce a single mean value. Then, the decrease of the cycle-averaged surface temperature as a function of the coolant temperature reduction is calculated and given
JER03005 IMechE 2005

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

Characterizing the thermal sensitivity of a gasoline HCCI engine

301

Fig. 12 Local wall temperature proles on piston and head surfaces measured for dierent levels of coolant temperature

(Fig. 14(a)). The local surface temperature variations follow the same trend, and their reduction is slightly larger than the coolant temperature drop (mean slope=1.186). This is as a result of a compounded eect of the slower combustion rates and lower peak gas temperatures observed for reduced coolant temperatures (Fig. 8(a)). However, the cylinder head and piston surfaces behave somewhat dierently. The head surface temperature is more directly driven by the coolant path through the cylinder head, and a slope of the line going through only cylinder head points is closer to unity. The piston surface temperature is more sensitive to variations in coolant temperature owing to greater sensitivity to the gas temperature histories varying with T (Fig. 8(a)). coolant In contrast, head surface temperatures increase by only 10 C when intake charge temperature varies
JER03005 IMechE 2005

from 75 C to 115 C under the constant coolant temperature of 95 C, as shown in Fig. 14(b). Thus, intake charge temperatures aect the core gas temperature much more than the near-wall conditions, and contribute much less to the thermal stratication than the variation of the wall temperature. Average piston temperature variation with intake charge temperature was not available.

5 COMPENSATING FOR THE WALL TEMPERATURE EFFECT BY INCREASING THE INTAKE CHARGE TEMPERATURE Experiments discussed in previous sections show that a reduction of the coolant temperature below 80 C, while keeping the intake charge temperature
Int. J. Engine Res. Vol. 6

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

302

J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt, and R Rask

Fig. 13 Local instantaneous heat uxes on piston and head surfaces obtained at dierent levels of coolant temperature

Fig. 14 Local mean surface temperature dierence due to the eect of (a) changing coolant temperature; (b) changing intake charge temperature
Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

Int. J. Engine Res. Vol. 6

JER03005 IMechE 2005

Characterizing the thermal sensitivity of a gasoline HCCI engine

303

constant, causes signicantly slower burning and reduces combustion eciency owing to the higher heat loss and partial burning in the expanded thermal boundary layer. Evaluations of heat ux trends purely as a result of wall temperature eects are, therefore, dicult, since the gas temperature histories vary from case to case. Hence, it is interesting to attempt to compensate for wall temperature eects with a simultaneous change of intake charge temperature in order to maintain constant combustion phasing and peak heat release rate. This would determine the required increment of intake temperature change for a given change of T , i.e. coolant allow direct comparisons of sensitivities, and provide quantitative insight into the thermal stratication between hotter gas core and cooler near-wall layer. The latter would be possible by eliminating the eect of burn rates on heat ux, thus allowing assessment of only the wall temperature eect on local heat ux. Specication of the test point is given in Table 3. Cycle-averaged piston and head surface temperatures are plotted in Figs 15(a) and (b). Wall temperatures on the cylinder head are changing linearly with the coolant temperature variation, while the

Table 3 Operating conditions with varying T and intake T dependently coolant


Fuel preparation
Fuelling rate (mg/cycle) Speed (r/min) Swirl control valve (deg) Air fuel ratio External EGR (%) Internal EGR (%) Intake manifold pressure (kPa) Exhaust manifold pressure (kPa) Control target for the location of peak pressure ( ATDC)

Fully premixed
11.0 2000 36 20.5 0 40 94.8 101.0 13.0

piston top temperatures show a slight deviation from a one-to-one correlation, presumably because of inconsistent oil temperature variation. It was not possible to reduce the oil temperature below 70 C because of test cell hardware limitations, as shown in Fig. 15(c). Spatial variations remain relatively the same throughout the range studied, as temperatures at individual piston and head locations follow the same trend. Figure 15(d) shows how much intake temperature needs to change to compensate for

Fig. 15 Mean temperature measurements illustrating a simultaneous variation of T (F) with intake T (E) in order to ensure constant burn rates: (a) average piston surface temperature; coolant (b) average head surface temperature; (c) oil temperature; (d) intake charge temperature
Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

JER03005 IMechE 2005

Int. J. Engine Res. Vol. 6

304

J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt, and R Rask

a given reduction of the wall temperature. The increment is linearly proportional to coolant temperature reduction, with a factor of proportionality of 1.11. A one-to-one proportionality line is included for reference. If only the extremes of the range are observed, a 55 C DT increment is required for intake the 45 C coolant temperature decrement. The temperature trends described above allowed maintaining constant burn-rate proles. The global combustion rate is unchanged, as indicated by the location of the 50 and 90 per cent burned mass points in Fig. 16. Figure 17 demonstrates that neither do emissions characteristics change much when the simultaneous change of T successfully compensates for the intake variation of T . coolant Despite the fact that bulk burn rates remain absolutely consistent over the range of T /T coolant intake combinations, as shown in Fig. 18(a), the spatially averaged heat ux proles given in Fig. 18(b) change systematically. Each of the proles in Fig. 18(b) represents an ensemble average performed over a total of nine locations (seven on the piston, two on the head). Ensemble averaging was chosen over areaweighted averaging because it was felt that physical conditions in the well-mixed HCCI engine for the

Fig. 16 Combustion phasing comparisons for a simultaneous variation of T (F) with T (E) intake coolant

Fig. 17 Emission comparisons for a simultaneous variation of T (F) with T (E) intake coolant
Int. J. Engine Res. Vol. 6

rst time create a situation where using ensemble averaged heat ux is justied, and accompanied with less uncertainty than performing area averaging in a chamber with four valves and a complex piston top shape. Local heat ux proles measured in an HCCI engine show very similar features, irrespective of the relative distance from cylinder axis or any other spatial reference, thus oering evidence that combustion in a premixed HCCI occurs simultaneously at multiple sites without any macroscopic ame propagation (see Fig. 13 and [11]). This is in great contrast to conventional SI and CI engines, where local variations are much more pronounced, owing to ame-front propagation and extremely heterogeneous composition in the chamber, respectively. The hypothesis that the spatially averaged heat ux can accurately represent the global process was checked in the previous study [11], by performing the complete heat release analysis using only measured quantities. Indeed, it was conrmed that adding measured heat loss (from integrated spatially averaged heat ux) to calculated net heat release derived from in-cylinder pressure trace satises energy balance for a range of conditions. Comparing average heat ux proles in Fig. 18(b) allows assessment of whether there is a global change in heat ux trends when wall temperature varies while burn rates remain constant. The peaks of spatially averaged heat uxes decrease and the tails extend longer as the wall temperature decreases. Phasing of combustion (location of the 50 per cent burned point) and peaks of the spatially averaged heat uxes are compared in Fig. 19(a). Combustion phase represents the global energy release rate, while spatially averaged heat ux phasing represents features of the averaged transient heat transfer rate from the boundary layer to the wall. The 50 per cent burn point occurs earlier than the location of the peak spatially averaged heat ux, indicating that the thermal boundary layer introduces a delay in transferring heat to the wall. For coolant temperature 95 C the phase dierence is 10 crank angle, and it gradually increases up to 17 crank angle as the coolant temperature decreases to 52 C. In Fig. 19(b), the location of 10 per cent burned is compared with the location of peak second derivative of heat ux. Previous study in a spark ignition engine [22] proposed that a ame arrival time at a certain location in the combustion chamber could be determined from the phasing of the second derivative of heat ux. For an HCCI engine, it could be interpreted as a start of energy release in the vicinity of the wall. In this sense, comparison in Fig. 19(b) illustrates the time dierence between the global ignition time
JER03005 IMechE 2005

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

Characterizing the thermal sensitivity of a gasoline HCCI engine

305

Fig. 18 Comparisons of measured (a) rate of heat release; (b) spatially averaged heat ux, during a simultaneous variation of T (F) with T (E) intake coolant

Fig. 19 Comparisons of (a) location of peak heat ux and 50 per cent mass fraction burned point; (b) location of the second derivative of heat ux and 10 per cent mass fraction burned point, obtained for a simultaneous variation of T (F) with T (E) intake coolant

and the beginning of local heat release near the wall. This delay is 0.25 ms (3 crank angle at 2000 r/min) at 95 C and it increases up to 0.375 ms (4.5 crank angle at 2000 r/min) owing to the thicker thermal boundary layer formation at lower wall temperature. A detailed insight into these phenomena can be gained from the analysis of local heat ux proles given in Fig. 20. These individual heat uxes, measured at dierent locations, were used for determining the spatially averaged heat ux shown in Fig. 18(b). Signicant variations of instantaneous local heat ux proles can be observed at some locations, while there seems to be no change at others. Locations number 1, 4, 5, and 6 show a systematic reduction of heat ux as coolant temperature decreases. In particular, locations number 5 and 6, located at the periphery and on the large piston squish area, display considerable heat ux decrease and retardation. This indicates that, while the global burn rates are equal (see Fig. 18(a)), the local burnJER03005 IMechE 2005

up rate at the periphery, particularly near the interface with the cylinder wall on the squish side of the chamber, decreases as the coolant temperature decreases. This can only be as a result of the thicker boundary layer and more mass being at temperatures suppressing the chemical reactions. The other four locations, three in the piston bowl (2, 3, 7) and one in the head (h1) do not show any signicant variations of the heat ux, thus the boundary layer at those locations does not vary greatly with T . coolant The probable reason is the squish ow being directed at those locations. Since the increased intake charge temperature raises the bulk gas temperature, it may be hypothesized that the hot gas being pushed out of the core towards locations 2, 3, and 7 counters the eect of near-wall thermal phenomena. In summary, enhanced thermal stratication (larger gradient of gas temperature distribution) owing to the colder wall surface enhances the spatial variations of the transient heat transfer rate. The local
Int. J. Engine Res. Vol. 6

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

306

J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt, and R Rask

Fig. 20 Instantaneous local heat uxes measured during a simultaneous variation of T (F) intake with T (E) coolant

heat ux measurements suggest that thickening of the boundary layer causes slower burning near the wall, particularly in the squish regions on the periphery of the piston, close to the interface with the liner. Lower surface temperatures and thermal stratication aect bulk burn rates signicantly owing to slower burning near the wall, therefore a relatively larger change of T was needed to comintake pensate for the eect of varying T . Somewhat coolant surprisingly, under conditions investigated in this study, the local stratication did not have a major
Int. J. Engine Res. Vol. 6

eect on the global emissions trends, but this would probably change under conditions closer to the misre limit.

6 CONCLUSIONS This paper focuses on the eects of the thermal environment on HCCI combustion and heat transfer. Two thermal parameters, i.e. intake charge temperature and wall temperature, were considered and their
JER03005 IMechE 2005

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

Characterizing the thermal sensitivity of a gasoline HCCI engine

307

eects on HCCI engine ignition and burn rates were characterized through systematic experimentation. The measurements were performed on a singlecylinder, four-valve, gasoline-fuelled engine with premixed charge and exhaust re-breathing. An additional series of tests was performed with a simultaneous variation of intake and coolant temperature in order to maintain unchanged burn rates by compensating for the eect of one with the change of the other. The results quantify the global thermal eects on ignition and combustion, but also provide insight into the fundamental dierences in the way the core gas temperature aects ignition/burning versus the thermal stratication and near-wall boundary layer phenomena. The studies of independent variations of intake or coolant temperature indicate the following. 1. Combustion in the tested gasoline HCCI engine has great sensitivity to thermal conditions. The peak burn rate decreases by ~35 per cent when the coolant temperature is reduced by only 15 C, while the relative change of peak burn rates corresponding to a 40 C reduction of intake temperature is more than 50 per cent. Thus, the wall temperature eect on HCCI combustion is stronger than the intake charge temperature eect. 2. The range of the coolant/wall temperature variations was severely limited at the low end by the partial burns and large cyclic variations. Cyclic variations prevented running the engine under 75 C coolant temperature. 3. The performance depends primarily on combustion phasing, hence excursions of either T intake or T away from the optimum reduced the coolant i.m.e.p. by up to 3 per cent at the low temperature limit. However, the temperature window for coolant is much smaller (15 C) than the intake charge temperature window (40 C). The emissions follow predictable trends, with NO decreasing x and HC and CO increasing with reduction of either T or T . coolant intake 4. Reducing the intake charge temperature retards ignition as a result of the lower core gas temperature and the burn rates decrease as a direct consequence of retarded ignition timing. 5. Decreasing coolant temperature indirectly aected charge temperatures and ignition timing, but it also reduced the burn rates much beyond what would be expected as a result of ignition timing retardation. The analysis of the correlation between the ignition timing and the burn duration
JER03005 IMechE 2005

suggests that reduced wall temperature extends the bulk burning because more charge is contained in the cooler near-wall thermal boundary layer, thus slowing down burn-up. 6. The highest surface temperature was measured at the centre of the piston top. The spatial variations of temperature were in the range of 10 C for the seven piston-top measurement locations. The magnitude of the dierence between the two locations at the head surface was also approximately 10. Local wall temperatures followed the change of T very closely, while the variations coolant resulting from T changes were minimal. intake 7. Measured local heat ux proles showed great sensitivity to varying wall temperatures, but in the case of independent variations of intake or coolant temperatures, most of the variations were driven by global burn rates. While local heat uxes remain quite uniform over the piston and head surface, the retardation of heat ux proles with reduction of T is somewhat more procoolant nounced for locations close to the periphery of the piston. The study of dependent variations of intake or coolant temperature, designed to maintain unchanged burn rates for any wall temperature and thus exclude the eect of burn rates on heat ux trends, showed the following. 1. Variations of combustion phasing and peak burn rates owing to wall temperature changes can be compensated for if the intake charge temperature is varied in the opposite direction and relatively more, with a factor of proportionality being 1.11. With such a test procedure, it was possible to vary the coolant temperature between 52 and 97 C and keep burn rates constant. 2. Emissions were not signicantly aected by wall temperatures in the case of simultaneous change of T . It appears that, as long as the bulk burn intake rates and combustion eciency can be sustained, the unburned HC and CO can remain unchanged. 3. Measured local heat uxes indicated distinctly dierent behaviour at locations closer to the periphery on the squish side of the asymmetric combustion chamber, than at locations on the opposite side. The locations in the squish zone demonstrated a systematic retardation of the heat ux prole with reduction of the wall temperature, particularly those close to the interface with the liner surface. This indicates a tangible growth of thermal boundary layers in the zone where the squish ow displaces the hot core gas and
Int. J. Engine Res. Vol. 6

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

308

J Chang, Z Filipi, D Assanis, T-W Kuo, P Najt, and R Rask

intensies heat loss to the wall. The local heat ux close to the centre and away from the squish zone remained virtually unchanged as the average wall temperature was reduced, presumably owing to the hot core gas being pushed by the squish ow towards these locations. 4. Spatially averaged heat ux prole showed dependence on the wall temperature, despite the consistent bulk burn rates ensured by a simultaneous change of the intake temperature. Consequently, the locations of the peak heat ux and the second derivative of heat ux, the latter indicating the start of vigorous burning in the vicinity of the thermocouple location, are retarded for reduced wall temperature. This conrms the hypothesis that cooler walls aect combustion primarily through the growth of the thermal boundary layer and slower burning near the wall. Overall, this work suggests a qualitatively dierent eect of intake charge temperature than the coolant temperature on HCCI combustion. While the intake charge temperature aects directly the thermal state of the bulk gas in the core, the coolant (wall) temperature variations produce a more signicant eect on the gas in the thermal boundary layer. Hence, the essence of the wall temperature eect is thermal stratication in the near-wall regions. This can be a signicant factor in controlling HCCI operation particularly close to misre or knock limits, hence motivating further research into precision thermal management.

ACKNOWLEDGEMENTS This research has been sponsored by the GM/UM Collaborative Research Laboratory for Engine Systems Research at the University of Michigan. The authors appreciate Orgun Guralps help with setting up the telemetry instrumentation and assisting in conducting measurements.

REFERENCES
1 Iida, M., Aroonsrisopon, T., Hayashi, M., Foster, D., and Martin, J. The eect of intake air temperature, compression ratio and coolant temperature on the start of heat release in a HCCI engine. SAE paper 2001-01-1880/4278, 2001. 2 Marriott, C. and Reitz, R. Experimental investigation of direct injection-gasoline for premixed compression ignited combustion phasing control. SAE paper 2002-01-0418, 2002.
Int. J. Engine Res. Vol. 6

3 Jun, D., Ishii, K., and Iida, N. Combustion analysis of natural gas in a four stroke HCCI engine using experiment and elementary reactions calculation. SAE paper 2003-01-1089, 2003. 4 Persson, H., Agrell, M., Olsson, J., Johansson, B., and Strom, H. The eect of intake temperature on HCCI operation using negative valve overlap. SAE paper 2004-01-0944, 2004. 5 Sjoberg, M. and Dec, J. An investigation of the relationship between measured intake temperature, BDC temperature, and combustion phasing for premixed and DI HCCI engines. SAE paper 2004-01-1900, 2004. 6 Sjoberg, M., Dec, J., Babajimopoulos, A., and Assanis, D. Comparing enhanced natural thermal stratication against retarded combustion phasing for smoothing of HCCI heat-release rates. SAE paper 2004-01-2994, 2004. 7 Dec, J. E. and Sjoberg, M. Isolating the eects of fuel chemistry on combustion phasing in an HCCI engine and the potential of fuel stratication for ignition control. SAE paper 2004-01-0557, 2004. 8 Press, W. H., Teukolsky, S. A., Vetterling, W. T., and Flannery, B. P., Numerical Recipes in C, 2nd edn, 1992, pp. 650655 (Cambridge University Press). 9 Sta, J. A universally applicable thermodynamic method for T.D.C. determination. SAE paper 200001-0561, 2000. 10 Brunt, M. F. and Pond, C. R. Evaluation of techniques for absolute cylinder pressure correction. SAE paper 970036, 1997. 11 Chang, J., Filipi, Z., Assanis, D., Kuo, T., Najt, P., and Rask, R. New heat transfer correlation for an HCCI engine derived from measurements of instantaneous surface heat ux. SAE paper 200401-2996, 2004. 12 Eichelberg, G. Some new investigations on old combustion engine problems. Engineering 1939, 148, 463446, 547560. 13 Assanis, D. N., Friedman, F. A., Wiese, K., Zaluzec, M. J., and Rigsbee, J. M. A prototype thin-lm thermocouple for transient heat transfer measurements in ceramic coated combustion chambers. SAE paper 900691, 1990. 14 Jackson, N. S., Pilley, A. D., and Owen, N. J. Instantaneous heat transfer in a highly rated DI truck engine. SAE paper 900692, 1990. 15 Borman, G. and Nishiwaki, K. Internal-combustion engine heat transfer. Prog. Energy Combust. Sci., 1987, 13, 146. 16 Assanis, D. N. and Badillo, E. Evaluation of alternative thermocouple designs for transient heat transfer measurements in metal and ceramic engine. SAE paper 890571, 1989. 17 Hayes, T. K., White, R. A., and Peters, J. E. Combustion chamber temperature and instantaneous local heat ux measurements in a spark ignition engine. SAE paper 930217, 1993. 18 Guralp, O. A. Development and application of a telemetry system for piston surface temperature
JER03005 IMechE 2005

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

Characterizing the thermal sensitivity of a gasoline HCCI engine

309

19

20 21

22

measurement in a homogeneous charge compression ignition engine. MS Thesis, Mechanical Engineering Department, University of Michigan, 2004. Opris, M. C., Jason, R. R., and Anderson, C. L. A comparison of time-averaged piston temperatures and surface heat ux between a direct-fuel injected and carbureted two-stroke engine. SAE paper 980763, 1998. Overbye, V. D., Bennethum, J. E., Uyehara, O. A., and Myers, P. S. Unsteady heat transfer in engines. SAE Transactions, 1961, 69, 461494. Chang, J. Thermal characterization and heat transfer study of a gasoline homogeneous charge compression ignition engine via measurements of instantaneous wall temperature and heat ux in the combustion chamber. PhD dissertation, Department of Mechanical Engineering, University of Michigan, 2004. Cho, K. Characterization of combustion and heat transfer in a direct injection spark ignition engine

through measurements of instantaneous combustion chamber surface temperature. PhD dissertation, University of Michigan, 2003.

APPENDIX Notation A ,B n n N k l a w n v Fourier coecient total number of harmonics thermal conductivity (W/mK) distance from surface junction thermocouple thermal diusivity (m2/s) harmonic phase angle (deg) angular velocity of engine crankshaft (rad/s)

Downloaded from jer.sagepub.com at INDIAN INST OF TECHNOLOGY on July 24, 2012

JER03005 IMechE 2005

Int. J. Engine Res. Vol. 6

You might also like