Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM SAE TECHNICAL PAPER SERIES 2007-01-1052 A Crevice Blow-by Model for a Rapid Compression Expansion Machine Used for Chemical Kinetic (HCCI) Studies S. Scott Goldsborough Marquette University Reprinted From: CI and SI Power Cylinder Systems, 2007 (SP-2073) 2007 World Congress Detroit, Michigan April 16-19, 2007 400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A. Tel: (724) 776-4841 Fax: (724) 776-0790 Web: www.sae.org Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM By mandate of the Engineering Meetings Board, this paper has been approved for SAE publication upon completion of a peer review process by a minimum of three (3) industry experts under the supervision of the session organizer. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without the prior written permission of SAE. For permission and licensing requests contact: SAE Permissions 400 Commonwealth Drive Warrendale, PA 15096-0001-USA Email: permissions@sae.org Fax: 724-776-3036 Tel: 724-772-4028 For multiple print copies contact: SAE Customer Service Tel: 877-606-7323 (inside USA and Canada) Tel: 724-776-4970 (outside USA) Fax: 724-776-0790 Email: CustomerService@sae.org ISSN 0148-7191 Copyright © 2007 SAE International Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE. The author is solely responsible for the content of the paper. A process is available by which discussions will be printed with the paper if it is published in SAE Transactions. Persons wishing to submit papers to be considered for presentation or publication by SAE should send the manuscript or a 300 word abstract of a proposed manuscript to: Secretary, Engineering Meetings Board, SAE. Printed in USA Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM 2007-01-1052 A Crevice Blow-by Model for a Rapid Compression Expansion Machine Used for Chemical Kinetic (HCCI) Studies S. Scott Goldsborough Marquette University Copyright © 2007 SAE International ABSTRACT A crevice blow-by model has been developed for a Rapid Compression Expansion Machine. This device can be used to study chemical kinetics with application to Homogeneous Charge Compression Ignition and other alternative combustion processes. In order to accurately resolve the ignition conditions and understand the oxidation process, accurate models for heat transfer and crevice flow, including blow-by past the ringpack, must be utilized. Crevice flows are important when high compression ratio or boosted operation is investigated. In previous work the heat loss characteristics of the RCEM were characterized; this study concerns the crevice flows within the RCEM. A ring-dynamic model, first developed at MIT and recently modified at UIUC to account for circumferential flow pas unlubricated rings, was employed. The 0-D model was coupled to a four-zone heat release code and tuned so that good agreement could be achieved between a computed ‘modified-entropy’ pressure and the experimentally measured pressure for a number of high compression ratio (20-50:1) ‘motored’ runs. The model was then used with two ‘fired’ cases (natural gas / air) to understand how the crevice dynamics change and how the model affects the understanding of the combustion process. In a companion paper the ringpack model is incorporated into an integrated chemical kinetics / computational fluid dynamics (CFD) code to investigate the influence of the crevice flows on the in-cylinder charge motion and temperature profiles that develop during the compression and expansion strokes of the RCEM. INTRODUCTION Rapid Compression Machines (RCMs) and Rapid Compression Expansion Machines (RCEMs) are often used to study low and intermediate temperature (<1100K) autoignition characteristics of fuel-air mixtures [1-24]. These devices are usually configured with piston-cylinder geometries where an initially quiescent fuel / air or air-only charge is rapidly compressed by the piston(s) to a desired temperature and pressure; compression times are on the order of 15-80ms. In this arrangement the combustion process can be isolated from typical in-cylinder engine phenomena like the turbulent fluid dynamics which result from the gas exchange processes. The chemical kinetics, which play a predominant role in homogenous charge compression ignition (HCCI) and many low temperature combustion (LTC) schemes, can thus be studied and evaluated. The data collected from these machines have been used to develop and correlate detailed and reduced kinetic models for various fuel species [25-33]. RCMs operate with a locking piston so that the compressed temperature / pressure environment can be sustained for long times (up to 130ms). ‘Ignition delay’ (IJD) is usually reported, where this is typically defined as the residence time required at the elevated temperature & pressure before autoignition occurs. The ignition time is often taken as the point of maximum pressure rise, or maximum light emission observed. RCEMs are a variant of the RCM where a non-locking piston is used. In these devices, the influence of the expansion process on the ignition and heat release events can be studied; the dynamic piston behavior more closely approximates the time-varying pressure and volume conditions found in IC engines. Even though an elevated temperature / pressure environment cannot be maintained with this design, and ignition delay times may not be directly comparable to RCM results, the adaptation can be useful since the ignition and combustion process, including the formation of pollutants, can be affected by the dynamics of the piston. Regimes of partial or late (after top dead center (TDC)) oxidation can be investigated. Additionally, useful characteristics such as rate of heat release (ROHR), mass burned fractions (F), and cycle thermal efficiencies (KTH) can be used as additional means to validate kinetic and other HCCI modeling techniques. RAPID COMPRESSION EXPANSION MACHINE A free-piston RCEM has been constructed by Sandia National Laboratories, California to investigate the autoignition characteristics of various fuel-air charges. The primary interest of previous studies was to assess the energy conversion efficiency and emissions potential of HCCI, with the intent to incorporate a rapid combustion system into a free piston based powerplant Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM [7]. The objective of the current work is to develop a framework to accurately determine additional parameters such as autoignition temperature (including temperature distributions within the device (see Ref. [34] which is a detailed kinetic / CFD study of the RCEM utilizing the 0-D crevice model described here)), ignition delay, ROHR and F of different fuel-air charges. The RCEM’s design and operation has been described in detail previously [7,35], however its principal operating characteristics are reviewed here. An illustration of the SNL RCEM is presented in Figure 1; its primary specifications are listed in Table 1. The RCEM consists of a double ended free piston that is enclosed in a double ended cylinder (there is a combustion chamber and a driver chamber on either end of the piston). The piston crown on either end is flat (as can be seen in Figure 2) creating a pancake chamber at TDC. Stainless steel caps are used for the cylinder heads, while cylinders of steel and cast iron have been used. The cylinder is wrapped with Briskheat electric heating tape and covered with Fiberfrax insulation, which is visible in the photo, to enable investigation of preheated charges (up to 75C). The SNL RCEM is pneumatically driven using high pressure helium (near 45-65MPa). The helium pressure is adjusted based on the compression heating desired; higher driving pressures yield greater compression of the cylinder charge. (RCM studies generally vary the diluent composition to achieve different compression temperatures since the compression ratio is not fixed. The gas is dumped to the driver (or back) side of the double ended piston where the volume is initially evacuated. A specially modified Nupro bellows valve allows rapid gas transfer for the driving process. The combustion side is initially charged with a homogeneous mixture of fuel and oxidizer / diluent (‘air’). (For some liquid fuels the combustion chamber can be filled with ‘air’ only (at atmospheric pressure) and then fuel is injected directly into the chamber using a stainless steel syringe. If this procedure is followed, time is allotted for the fuel to evaporate and mix with the air, usually on the order of 30 minutes.) The fuel / air charges, whether premixed or directly injected, are assumed to be uniform and quiescent at the start of the compression process. The cylinder pressure is recorded using two piezoelectric effect transducers (Kistler types 7061A, 7063A, 7061B, 607L, and AVL Type QC42D-X with Kistler Type 5010 and 5026 charge amplifiers have been used) at a sampling frequency of 500kHz. The transducers are coated with a 0.25mm-thick layer of Silastic J silicone compound to prevent against thermal shock during the fired runs. The piston position is recorded using a Data Instruments FASTAR Model FS5000HP inductive transducer with a sampling frequency of 200kHz. For the data presented here the pressure traces were filtered using an FFT low-pass filter where the cutoff frequency was set to 4kHz for the ‘motored’ (air-only) runs and 30kHz for the ‘fired’ runs. Lower cutoff frequencies tend to significantly alter the Figure 1 – SNL Rapid Compression Expansion Machine. Bore 76.2 mm Stroke 254 mm (max) 234 mm (average) Bore to Stroke Ratio @ TDC 5.1 @ CR = 16:1 9.6 @ CR = 30:1 16.9 @ CR = 50:1 Piston Mass 2.89 kg Compression Time 13 - 15 ms Expansion Time 10 - 12 ms Compression Ratio Variable; utilized 16-50:1 Pressure Capacity >30MPa Table 1 – SNL RCEM Specifications perceived ignition point and the ROHR calculations (as illustrated in Ref. [35]). Noise from the position trace was removed by the following method. The position data was first smoothed using a binomial routine with the number of smoothing passes set to 200; the trace was then differentiated to obtain the piston velocity. The velocity trace was smoothed, again using the binomial routine (this time the number of passes was set to 1000); the trace was then differentiated to obtain the piston acceleration. The acceleration trace was smoothed, again using the binomial routine (this time the number of passes was set to 6000). The velocity trace was then integrated to recover the piston position. Examples of this procedure and the effects that this can have on the data can be found in Ref. [35]. The SNL RCEM has been designed to operate at high pressures, measured up to 30MPa; this design parameter is important since some charges, such as lean natural gas / air mixtures shown here, can require significant compression heating (up to 45:1, 17MPa) before the autoignition process is initiated. Additionally, accurate ignition/oxidation data is needed for high pressure environments. In order to achieve this level of pressurization with minimal piston and head blow-by, a custom ringpack and head sealing system have been designed. Three ringpack configurations have been utilized, each using different in ring types and groove geometries; for more detail see Ref. [7]. One Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM Combustion End Driver End Figure 2 – RCEM free piston and detailed ringpack. configuration, which was used in this study, is illustrated in Figure 2. This ringpack consists of two seals on the combustion chamber end and one on the driver end. The combustion or top-land seal is a C. Lee Cook ZGS compression ring (graphite reinforced polyimide), and behind that a Furon elastomeric lip seal (lubricated glass-filled PTFE) with an o-ring expander ring (this ring is a continuous piece). On the driver end a Furon lip seal is used in conjunction with a stainless steel expander spring. Two C. Lee Cook bronze-impregnated Teflon rider rings are used to ease the piston’s motion within the cylinder. The head uses a plain o-ring which is fitted into a rectangular cross-section groove cut into the cylinder body. A 45˚ bevel is cut around the circumference of the interior cylinder and the piston crown to enable easy insertion of the free piston into the cylinder. The head is fastened to the cylinder body using 12 bolts positioned around the circumference. AUTOIGNITION PARAMETERS RCM and RCEM data sets can include autoignition temperatures, ignition delays, rates of heat release and mass burned fractions; sometimes specie concentrations (temporally varying, or at a single time) are available. However, in order to correctly characterize the data sets an accurate quantification of the in-cylinder temperatures through the compression and delay, or expansion processes is required; temperature is the predominant driver of the chemical kinetics and thus it is extremely important to accurately determine this. Various experimental and computational techniques have been employed to evaluate this parameter, including the use of fine-gage thermocouples [18], planar laser induced fluorescence [35,37], Rayleigh scattering [38], 1-D gas dynamic calculations [39] and multi-dimensional computational fluid dynamics software [38,40,41]. Recent RCM designs have incorporated features that seek to minimize the thermal gradients within the charge; the objective is to create a uniform charge with a single representative temperature. This is important with regard to interpreting the rates of reaction, especially in regimes where significant negative temperature coefficient (NTC) behavior can be observed. In addition, the ability to use 0-dimensional Figure 3 – RCM piston with machined crevice volume [46]. Figure 4 – RCF piston utilizing a nose-cone geometry [18]. (0-D) modeling of the system is preferred due to lower computational costs. Finally, substantial stratification within the reaction chamber and the resulting gradients in the heat release time can lead to the development of resonant acoustic waves which can alter the heat loss characteristics of the charge and emissions formation within the system. These are not indicative of the bulk gas kinetics themselves and should be avoided. Modern RCM designs have utilized machined piston crevice volumes or nose-cone piston geometries [18,38,41,42] (for examples see Figures 3 and 4) which work to swallow, or exclude the cold gases that are adjacent to the cylinder walls during compression, and prevent their reintroduction into the bulk charge during the delay period; the vortex roll-up behavior that is usually seen in reciprocating engines [43,44] is suppressed or excluded entirely. O-rings are typically used to seal the reaction chamber during the experiment. RCEMs are precluded from using significant crevice volumes as this can enable large quantities of cold gas to reemerge into the bulk charge on the expansion stroke; crevice flows should thus be minimized. In some studies the piston(s) have been coated with silicon-based thermal barriers to reduce the heat losses that result from the “vortex roll-up” and thus decrease the degree of thermal stratification that develops during compression [7,35]. Engine-like ringpacks are often used in RCEMs to seal the reaction chamber [35,45]. The bulk charge temperature in RCMs is typically computed using a 0-D, isentropic compression approximation; variations in specific heats are Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM considered and either the geometric compression ratio or the experimental pressure ratio used. Heat loss has been incorporated into the analysis through the use of either a time-varying cylinder volume [24,31,38], an effective compression pressure [18,19,22], or a global heat transfer correlation [28,29]. The charge temperature in RCEMs has been computed using an equation of state (EOS) (e.g., ideal gas (similar to IC engine analyses)); a ‘modified-entropy’ approach has also been used in conjunction with a global heat transfer correlation [35,45,47]. Another method is to replace the isentropic coefficient with a polytropic coefficient. Mass transfer to and from the crevice region, including blow-by past the rings, can significantly alter the experimental pressures achieved, as well as the incylinder temperature distributions, and the emissions recorded during the experiment. This is especially true if high compression ratios are utilized since a greater fraction of the bulk charge can enter and leave the crevice volume; boosted charge operation can have a similar effect. Calculations of the ROHR can also be skewed if the crevice flows are substantial but not taken into account. Because of these issues it is important to appropriately quantify the crevice flows for the RCEM studies; this is the primary objective of this paper. The remainder of this manuscript is organized as follows. First an overview is presented of the two prominent crevice modeling techniques; this is followed by a detailed description of the ring-dynamic model explored in this study. The specifics of the bulk charge modeling are presented along with some notes concerning the solution methodology. Finally, results of the RCEM modeling are presented, illustrating both ‘motored’ and ‘fired’ conditions, with conclusions drawn concerning the significance of the crevice flows in the RCEM. CREVICE FLOWS Two approaches are typically utilized to account for crevice mass transfer in piston-cylinder arrangements. The first, a relatively simple method, assumes a single aggregate, fixed-volume to which all of the crevice charge is transferred; this is assumed to be in pressure equilibrium with the main cylinder gases, and in thermal equilibrium with the piston and cylinder wall. Surface area to volume ratios are generally large enough to satisfy this latter condition, enabling adequate rates of heat transfer between the crevice gases and the solid engine surfaces to achieve isothermality. (This was demonstrated by Furuhama and Tada [48] to be a good approximation in operating IC engines.) This volume could be considered as the top land crevice volume, that region just above the top compression ring. Calculating mass flow to and from this region using this method is fairly straightforward and computationally inexpensive; as such, it is often used for real-time engine ROHR calculations [49,50]. A second, more complex, and more computationally expensive approach employs a series of disparate volumes which represent the various ‘pockets’ within a piston ringpack. The pressures of these volumes are not assumed to be in equilibrium with the main cylinder gases, but are based on the time varying flow rates into and out of these ‘pockets’ and the size of the ‘pockets’. In this model, as with the former, thermal equilibrium with the solid surfaces is usually assumed. This ‘ringdynamic’ crevice model, as it is termed since it also accounts for the dynamic movement of the rings within the grooves of the ringpack, was initially proposed by Namazian [51] and Namazian and Heywood [52]. This model has been used for analyses of emissions and the influence of the crevice flows on the in-cylinder fluid dynamics [51-59]. Figures 5a & 5b are presented to illustrate the differences between these two approaches, as applied to the SNL RCEM. In these figures computed and measured pressure traces, and computed temperature traces are shown for a ‘motored’ shot, versus instantaneous compression ratio. Much of the data in this work are plotted versus instantaneous geometric compression ratio (i.e., Vmax / Vinst) to eliminate the slight variations in piston compression - expansion times (for varying driver pressures, i.e., maximum compression ratios), and to accentuate the region near TDC, where the largest rates of mass transfer occur. Arrows indicate the direction of the traces; here along a clockwise path. For this run the cylinder contained air only and a maximum geometric compression ratio of about 50:1 was used. The initial pressure and temperature were 0.099MPa and 300K, respectively. The computed pressures are calculated using a ‘modified-entropy’ approach where the cylinder charge is assumed to be compressed and expanded with constant specific entropy (between each time step (i.e., the intervals of piston motion data)) with the specific entropy modified to account for heat losses. Details of this procedure can be found in the Bulk Charge Model section of this paper. The crevice model parameters have been adjusted to achieve fairly good agreement with the peak pressures, and, for the ring-dynamic model, with the cylinder pressure along the piston’s trajectory. For this comparison, and throughout the rest of this paper, the combustion chamber volume is computed by assuming a right-circular cylinder configuration defined by the bore and the measured stroke; a ‘free-volume’ is added to account for slight variations that may occur as the piston position is initialized. ‘Free-volume’ values were within r2.0cm3 (about 0.2% of the initial cylinder volume); this parameter was adjusted for each run in order to enable good agreement between the measured pressure trace and the ‘modified-entropy’ pressure. As can be seen in the figures, the ring-dynamic model gives significantly better agreement with the experimentally measured pressure and computed massaverage temperature both on the compression stroke and the expansion stroke. In Fig. 5b it can be seen that Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM 18 Pressure [MPa] Air-only Charge Compression Ratio 50:1 Initial Temperature 25C Initial Pressure 0.099MPa Significant discrepancy through compression and expansion strokes for single zone model 12 Slight departure on expansion stroke for ring dynamic model 6 a1_17 Experimental pressure Computed pressure, ring dynamic model Computed pressure, single zone model 0 0 10 20 30 40 50 Instantaneous Compression Ratio Figure 5a – Cylinder pressure vs. instantaneous compression ratio; comparison of experimental data to calculations using singlevolume crevice model and ring-dynamic model, with maximum geometric compression ratio of 50:1. 1200 Temperature [K] Air-only Charge Compression Ratio 50:1 Initial Temperature 25C Initial Pressure 0.099MPa 900 Slight departure on expansion stroke for ring dynamic model Significant discrepancy through compression and expansion strokes for single zone model 600 Modified-entropy Temperature Mass averaged temperature, ring dynamic model Mass average temperature, single zone model a1_17 300 0 10 20 30 40 50 Instantaneous Compression Ratio Figure 5b – Computed temperature vs. instantaneous compression ratio; comparison of modified-entropy temperature to mass-average temperature computed using single-volume crevice model and ring-dynamic model; a maximum geometric compression ratio of 50:1 is used. the mass-average temperature using the single-zone crevice model is substantially greater than the ‘modifiedentropy’ temperature, which is not realistic; this would indicate greater-than isentropic compression. (The mass-average temperature is computed using an EOS with the measured pressure, and the cylinder masses determined from the associated crevice model, e.g., Tavg = f (Pavg, ȣ); ȣ = Vcyl /(mcyl R).) Because of the better agreement achieved with the ring-dynamic model, this is the focus of this paper. RING-DYNAMIC MODEL A description of the ring-dynamic model is presented next. The ring-dynamic model attempts to account for the pressurizing and depressurizing of the ringpack pockets, the mass transfers from the main cylinder and between these pockets, the variations in the forces acting on the rings, and the kinematics of the rings within the grooves, with this generally limited to axiallyconstrained movements. Namazian and Heywood [52] coupled this 0-D model to experimental, in-cylinder pressure measurements to investigate the influence of the ringpack parameters on a number of factors, including the mass entering the crevice, the mass trapped within the ringpack, the mass lost to blow-by, and the contribution of the ringpack gas flow to the engine-out emissions. Kuo et al [53] coupled this model to a ring-friction model which interactively computed the lube-oil-film thickness; they conducted validation tests over a range of operating conditions and performed a sensitivity study of the model’s parameters. Reitz and Kuo [54] integrated the 0-D ring-dynamic model into a computational fluid dynamics (CFD) code (KIVA [55]), assuming uniform, axisymmetric behavior of the crevice flows, in order to investigate the influence of this gas motion on the bulk-charge fluid dynamics, as well as the in-cylinder temperature fields and the engine-out emissions for three different operating points of a sparkignited (SI) engine. Tonse [56] integrated the ringdynamic model into KIVA as well, also assuming axisymmetric behavior, in order to study the crevice flow induced fluid dynamics and the effects on a sparkignited propagating flame. Huynh et al [57] used a coupled 0-D / CFD (KIVA) formulation to study the influence of various engine operating parameters on the crevice flows and the resulting hydrocarbon emissions. Roberts and Matthews [58] modified the original formulation of the 0-D model to account for thermal expansion of the rings and piston, variations in gas composition within the ringpack, as well as azimuthal variations in the crevice flows, which result from the location of the ring end gaps. They compared calculations with spatially resolved, in-cylinder HC measurements, and investigated the influence of a number of engine operating parameters on the HC emissions. More recently, Zhao and Lee [59] modified the original ring-dynamic model to account for circumferential flows that occur in unlubricated engine configurations; their application was a small-bore directinjection (DI) optical engine. They integrated the 0-D model into KIVA to investigate the crevice flow induced, in-cylinder fluid dynamics and the blow-by emissions encountered in their engine. Finally, Potokar and Goldsborough [34] incorporated the 0-D model developed in this paper into an integrated chemical kinetic / CFD solver to investigate charge motion and temperature profiles within an RCEM. The ring-dynamic crevice model is presented here in two parts: first the mass transfer to and from the crevice and between the ringpack pockets is described, then the movement of the rings within the ringpack is discussed. The formulation used in this study is similar to the development by Namazian and Heywood [52], along with the modifications of Zhao and Lee [59]; additional details can be found in those references. Differences with previous approaches will be highlighted. Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM dm 3 dt dm 4 dt 13 m 23 m 13 c m 34 m 35 m 3 5c m (2) 34 m 45 m (3) In these expressions the mass flow rates between zones i j ; the “c” ‘i’ and ‘j’ are designated by subscripts, e.g. m subscript denotes the circumferential flow path. Figure 6 – Pathways for gas motion within the ringpack. GAS MOTION The ringpack pockets are pressurized and depressurized as mass travels from the bulk charge into the top-land crevice and then through various pathways within the pack. Due to the small size of the pathways the flow is restricted, and this results in a lag between the pressurization/depressurization of the main charge and the changing pressure of the ringpack gas. The primary passages for the gas motion are illustrated in Figure 6; these include the ringside clearance (between the inner diameter of the ring and the grooves of the piston), the end gaps of the rings (used for ring installation and removal), and the circumferential passage (between the outer ring diameter and the cylinder wall). The pockets within the ringpack are labeled 1 through 5 in the figure. For this work, these volumes are assumed to be discrete but azimuthally contiguous within the ringpack, with each zone independently homogeneous in temperature, pressure and composition. The gas in volume 1, the top-land crevice, is assumed to be at the cylinder pressure, which is consistent with previous studies. The gas in volume 5 is assumed to be at atmospheric pressure, which is consistent with previous work. As well, the gas within the pockets is assumed to quickly equilibrate to the wall temperatures. However, the effects of ring heating (due to sliding friction) on the flow rate and temperature of the charge passing through the circumferential gaps are taken into account. A discussion of the ring heating is presented in the Ring Heating sub-section. The mass conservation for each pocket can be expressed as in Equations (1) through (3). Previous publications have reduced these equations to rates of pressure change by assuming isothermal, ideal gas behavior within the pockets. In this work the expressions are kept in their generalized form so that an arbitrary EOS can be applied with the intent to account the effects of non-ideal gas behavior. (This will be more fully investigated in a future publication.) dm 2 dt 12 m 23 m (1) To determine the mass flow rates the flows between the zones are modeled as either channel (Poiseuille) flow, or orifice (isentropic) flow; these are described next. The flows into and out of volumes 2 and 4, the ringside clearance volumes, are modeled as compressible, laminar channel flows; the channel clearance heights are ct and the channel lengths (i.e., the ring radial lengths) are L. The flow rates using these assumptions can be determined by m A ct 2 U dP 12 P dx (4) With an assumed linear drop in pressure across the channel length this can be rewritten as m A ct 2 1 u dP U dx 12 P L ³ d dx (5) In this equation the upstream and downstream conditions are denoted by the limits ‘u’ and ‘d’; A is the cross-sectional area normal to the flow based on the clearance height and the circumference of the ring. The gas viscosity, µ, is assumed to be only a weak function of the pressure drop, and thus it is left outside the integral. The gas viscosity is evaluated at an average temperature, between entering and exiting conditions, where the formulation P CP T 0.7 (6) is used. The constant CP is set to 3.3e-7 kg/m-s, as per the discussion in Ref. [60]; this equation, though it was developed for hydrocarbon/air charges also fits data for dry air very well over the temperature range considered. The second set of flows, those through the ring end gaps, can be modeled as compressible, laminar, isentropic flow through an orifice. The 1-D flow rate can be written as m A Cd Us v s Cd Us 2 h o h s (7) where Cd represents the discharge coefficient, Us the downstream density (after the isentropic expansion), and Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM vs the throat velocity of the orifice. The velocity is rewritten in terms of the upstream, or stagnation enthalpy, ho, and the downstream, or throat (isentropic) enthalpy, hs, based on the isentropic condition. If the flow is choked the throat velocity approaches the sonic velocity where this can be expressed as, c § wP · ¨ ¸ © w U ¹s (8) The sonic velocity, c, at a particular pressure and temperature can be determined through an arbitrary EOS. (For ideal gas, this reduces to J R T .) It is noted here that previous publications used the ideal gas approximation to reduce Eq. (6) to the following, m A ­ 2 ª § P ·2 ° «¨ d ¸ Cd Uu c ® °¯ J 1 «¬ © Pu ¹ J §P · ¨ d¸ © Pu ¹ J 1 J 1 2 º ½° »¾ »° ¼¿ (9) Figure 7 – Free body diagram of the compression ring within the ringpack; the elastomeric ring is identical. of motion to yield the second derivative of the ring’s axial position within the ring groove; this is an inertial reference frame fixed to the piston. This is the same as the second derivative of the ringside clearance height can thus be expressed as, m r acr mr d 2 zcr dt 2 mr d 2 ct dt 2 FP Ff Fi Foil (10) where ȡ is the upstream density, Ȗ the ratio of specific heats (assumed constant), and the subscripts ‘u’ and ‘d’ represent the upstream and downstream conditions. However, references [52,53,59] have mistakenly written this expression without the square root of the 2/(Ȗ-1) term. In this equation the subscript ‘r’ denotes the ring, ‘p’ pressure, ‘f’ friction and ‘i’ inertia; the variable h is the clearance height, as in Eq. (5). These forces are determined using the following equations. The pressure force can be computed using the pocket pressures, The third set of flows, those through gaps between the cylinder wall and the outer ring surface (the circumferential flow that occurs due to the absence of a lubricating oil) are also modeled as compressible, laminar, channel flows, as per the discussion given in Ref [59]. Eq. (5) is used where it is also assumed that the temperature gain due to ring heating is linear along the length of the channel. The gas is heated to an average of the ring surface temperature and the temperature of the cylinder wall. FP1 A r1 P P3 P1 P2 A r1 2 2 2 (11) where the numeric subscripts denote the ringpack volumes; Ar1 is the ring-side surface area. The friction force as the ring slides along the cylinder wall is determined by Ff1 C f1 P2 A f1 C f1 P2 S d r1 t r1 (12) where C f1 is the coefficient of sliding friction. Again, the RING MOTION In a typical ringpack, the ring dynamics can contribute significantly to the flow through the crevice region, accounting for as much at 50% of the blow-by mass [51]. As the rings shift position within the grooves the pathways for the flow through the ringpack change and as a result the cylinder charge can more easily escape through the ringside clearance volumes, 2 and 4, from volumes 1 to 3, and 3 to 5. The kinematics of the rings are generally governed by a number of forces including contributions due to pressure, friction, inertia and the oil film; in the RCEM however, there is no lubricating fluid so the film force is absent. A free body diagram of the rings and their forces is presented in Figure 7; only axial dynamics are considered here, which is consistent with previous work. The forces shown can be combined using Newton’s law friction does not consider any twisting of the rings which could be important. Af1 is the surface area of the ring in sliding contact with the cylinder wall; this is given by the outer diameter of the ring and its axial length. The inertial force is the mass of the ring, mr, multiplied by its acceleration, ar, as in Fi mr ar (13) The acceleration of the ring in the non-inertial reference frame is generally governed by the piston motion, excepts when it floats freely within the piston groove, and for the purposes of this model it is assumed to be accurately approximated by the acceleration of the piston. (As is seen later in Figure 11, the inertial force is extremely small relative to the pressure and friction forces, and therefore any errors associated with ring acceleration term are negligible.) Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM BULK CHARGE MODEL – HEAT RELEASE ANALYSIS Figure 8 – Schematic of ring material for unsteady energy equation. RING HEATING As the piston traverses the cylinder length in the RCEM, the rings slide along the unlubricated wall. Due to the single-shot nature of the RCEM, some amount of transient friction heating is expected to occur. This effect may be more pronounced if rings of low thermal diffusivity are used. In order to consider the ring heating due to friction, especially at the ring surface, and any effect this might have on the flow rates or emerging gas enthalpies, the time varying temperature distribution within the ring is computed by applying the onedimensional, unsteady heat diffusion equation to the solid ring material. This is expressed as w Tr wt Dr w 2 Tr w x2 (14) and is assumed to be constant with respect to temperature. The heat flux into the ring surface due to sliding friction can be calculated by f fric Ff1 A f1 vr The methodology used here is multi-zonal: within the cylinder the charge is assumed to consist of three zones: a fresh zone, a burned zone, and a mixed zone; a crevice zone exists separate from the main charge and the mixed zone contains fresh and burned gases that reemerge from the crevice. The cylinder zones are assumed to be compressed through successive time steps with constant specific entropy and the partial pressure of each zone in the main charge is summed to give the total cylinder pressure. The cylinder pressure can be represented by Pcalc where Tr is the ring temperature, t is time, and x the radial direction into the ring, as seen in Figure 8. A cartesean formulation is used here as an approximation to the cylindrical arrangement since the diameter of the rings is large compared to the depth of heat penetration. The thermal diffusivity of the ring is represented by D r , qccf1 The ring-dynamic crevice model is integrated with a model for the bulk cylinder charge so that the effects of the crevice flows on the cylinder dynamics can be characterized. The model for the bulk charge is based on an energy balance approach that has been used extensively for the analysis of heat release in IC engines. This approach was originally described by Kreiger and Borman [61], Gatowski et al [62], Heywood [63] and Heywood and Chun [64] for SI engines, and is similar to more recent work by Jensen and Schrammof [65], and others [66-70]. (15) where f fric is the fraction of the friction power that is absorbed by the ring; the remainder of this friction heating should be absorbed the cylinder wall. Since the rings travel along the length of the cylinder wall, and because the cylinder wall has a relatively high thermal diffusivity, it is expected that there is no appreciable rise in the surface temperature of the wall. For the calculations presented here, the friction fraction is assumed to be 0.35. Larger values (to 0.5) give unrealistically high peak surface temperatures; smaller values (to 0.0) do not allow the computed bulk cylinder pressure to be accurately matched to the experimental pressure (see Fig. 13, and its discussion). Pfr Pbr Pmx (16) where the subscripts ‘fr’, ‘br’ and ‘mx’ indicate the fresh, burned and mixed zones, respectively. Pcalc is the total calculated cylinder pressure. The mass of each zone is allowed to vary through the compression / combustion (for a ‘fired’ run) / expansion processes as mass is ‘transferred’ between the cylinder and the crevice zone; for the ‘motored’ calculations there is no burned mass. In general, the fresh, burned and mixed charges are all allowed to enter the crevice, with the fraction from each based on the mass fraction of the zones within the cylinder. While it may be true that more of the unburned charge lies near the cold cylinder walls, and mass from this zone is more likely to enter the crevice, this assumption is used here anyway; this approach may need to be revisited for cases where slow burning fuel charges are investigated. The zones are assumed to be homogeneous and no inter-zonal heat transfer is allowed, however heat loss to the combustion chamber walls is incorporated into the analysis by modifying the zonal specific entropy at each time step, as dictated by the second law of thermodynamics (thus the term, ‘modified-entropy’ approach). For this work the total heat loss to the cylinder walls is calculated based on a modified Woschni model, which was correlated to the RCEM heat flux data [35]; the zonal heat loss is apportioned based on each zone’s representative volume fraction. The zonal specific entropy at a particular time step can be calculated as Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM 1 s fr ,t ' t m fr ,t ' t m ª¬ m fr s fr m fr , i s fr ,i m fr , e s fr , e q w Aw T fr m fr T fr m br Tbr m mx Tmx 0 The heat loss from the fresh zone is approximated by w Aw V fr fr V cyl q w Aw 3.82 P 0.8 ª 2.28S p 3.61×10 -3 0.73 ¬ H 0.2 Tcalc V ref Tini Pini Vini P calc Pmotor 0.8 º¼ T calc Tw m fr u fr 0 (21) 0 where the burned specific internal energy, ubr is a function of the zonal temperature, and pressure, generally. The integrated heat loss to the walls, the boundary work and total heat lost within the crevice volumes, Qcr, are included in this expression. The total crevice heat loss is calculated by applying the energy conservation equation to the crevice volume, Q cr m cr u cr m t cr u cr 0 ³ t 0 cr,ex h ex dt m t (22) cr ,in h in dt ³ m 0 where the integrated enthalpy flows into and out of the crevice used. (19) This formulation differs from previous heat release analyses due to the generalized equation of state employed. The advantage however, is that the effects of non-ideal gas behavior, which may be important in the crevice volume and in the main bulk charge for boosted or high compression ratio, low temperature combustion (LTC) operating schemes, can be taken into account. (20) SOLUTION METHODOLOGY For reference, the modified Woschni model is q w t t t dt Q ³ q w A w dt ³ W cr (18) m fr m br m mx m br u br m mx u mx m cr u cr @t 't zonal temperature is T. The zone-averaged, or ‘modified-entropy’ temperature is Tcalc, where this is computed by q u fr m br u br m mx u mx m cr u cr (17) fr This expression describes the fresh zone, as noted by the subscript, ‘fr’. The subscripts ‘t’ and ‘t+ǻt’ indicate the state of the system at different times separated by ǻt. The specific entropy is given by s, the rate of heat transfer to the wall by q w and the wall area by Aw. The Tcalc fr where the leading coefficient and the coefficient corresponding to the combustion generated portion (the second term in the brackets) have been changed from the original formulation [71] to match the RCEM data. In Eq. (20) H is the instantaneous height of the combustion chamber when this dimension is less than cylinder radius (otherwise this characteristic dimension is the bore), and the subscript ‘ini’ refers to the cylinder conditions at the beginning of the single-shot experiment. Pmotor for this analysis refers to an uncombusted pressure trace, taking into account heat and mass losses. Tw is the wall temperature. For the ‘fired’ runs the mass fractions burned are determined by matching the computed pressure as expressed in Eq. (16), to the measured pressure. It is assumed that any departure from the modified-entropy, or ‘polytropically’ computed pressure, after already accounting for heat transfer to the walls and mass transfer to or from the crevice, must come from a heat releasing reaction converting fresh or mixed charge to burned charge. The temperature of the burned zone, which is assumed to be of frozen composition (complete combustion with no dissociation), is determined by applying the energy conservation equation, written as, Some notes regarding the solution methodology are discussed here; details concerning the fitting of the crevice model parameters are provided in the next section and in the Appendix. The crevice and bulk charge models presented earlier in this paper are formulated in a generalized form; an arbitrary equation of state can be applied in their numerical solution. However, some of the calculations will require an iterative loop (for instance, using a Newton method to determine the sonic velocity at a particular pressure and temperature) as opposed to a simple algebraic equation. This results in a computationally more intensive solution. A forthcoming article will provide additional details of the methodology, and to illustrate the notable differences between the application of real gas equations of state and the ideal gas approximation. For the results presented here however, the RCEM dynamics are solved using the ideal gas EOS. The surface temperatures of the rings in the ringpack are computed using the unsteady energy equation (Eq. (14)) with a 4000 node mesh, just to the inside diameter of the rings. An explicit formulation is used with the node spacing chosen to satisfy the Fourier number criterion. The heat flux due to the friction is assumed to be uniform across the ring surface and a constant Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM temperature condition is applied to the inner boundary of the ring. Topland crevice height 0.254 cm Topland crevice volume (vol1) 0.100 cm3 The trapped ringpack charges, especially the one in volume 3, are all assumed to quickly lose any heating/cooling effects from the gas flow process and thus achieve isothermality within the pockets between time steps. The exception to this is the reemerging circumferential flow (from volumes 3 to 1), where the heat absorbed from the circumferential ring is carried through to the bulk charge. Volume behind compression ring (vol2) 0.040 cm3 Volume between the rings (vol3) 0.500 cm3 Volume behind elastomeric ring (vol4) 0.420 cm3 Compression ring thickness (t r1) 0.437 cm Compression ring width (rw1) 0.478 cm Compression ring mass (m r1) 11.23 g Compression ring ring-gap clearance (A 13) 0.000 mm2 Compression ring clearance height (ct1) 0.005 mm Compression ring circumferential area (A 13c) 2.500 mm2 Compression ring friction coefficient [72] 0.24 Compression ring density [73] 1412 kg/m3 Compression ring thermal conductivity [73] 0.536 W/m-K Compression ring heat capacity [73] 1100 J/kg-K Elastomeric ring thickness (t r2) 0.495 cm Elastomeric ring width (rw2) 0.508 cm Elastomeric ring mass (m r2) 6.09 g Elastomeric ring ring-gap clearance (A 35) 0.000 mm2 Elastomeric ring clearance height (ct2) 0.700 mm Elastomeric ring circumferential area (A 35c) 1.0-1.5 mm2 RCEM MODELING / ANALYSIS The goals of this modeling and analysis effort were to first demonstrate that the integrated crevice and bulk charge models can accurately interpret the RCEM experimental data, using the bulk charge parameter of average cylinder pressure for comparison. The effects of various modeling parameters (e.g., the accounting for ring heating, etc) on the computed results were also of interest. In addition, some understanding was to be gained with regard to the crevice flows, and the effects these may have on accurately quantifying the autoignition data, including the average cylinder temperature, the ignition delay, rate of heat release and mass burned fraction. The RCEM data set considered for this work included air-only ‘motored’ runs and natural gas / air ‘fired’ runs. The air-only experiments utilized geometric compression ratios ranging from 25-50:1, where the initial temperature and pressure of the charges were 25C and 0.099MPa, respectively. The natural gas / air shots were compressed to about 40:1 from an initial temperature and pressure of 67C and 0.1MPa, respectively. The natural gas consisted of 93.13% methane, 3.2% ethane, 0.7% propane, 0.4% butane, 1.2% carbon dioxide, and 1.37% nitrogen by volume. Driving pressures near 60MPa were utilized. Table 2 lists the parameters for the ring-dynamic model. The single-zone model has a single parameter, the volume of the zone; this was set to 0.60cm3 to match the experimental peak pressure of a 50:1 shot (see Fig. 2). The ring-dynamic model parameters are based on geometric values of the RCEM ringpack, however they have been modified to give reasonable agreement between the computed and experimental pressure profiles. Discussions of the ‘motored’ runs are presented first; the results of the ‘fired’ runs are presented after that. A sensitivity analysis of the model parameters is included in the Appendix for reference. ‘MOTORED’ RUNS Figure 9 illustrates the effectiveness of the ring-dynamic model towards accurately matching the experimental RCEM pressure; the major energy flows seem to be properly taken into account. In this figure the ratio of the experimental pressure to the ‘modified-entropy’ pressure is plotted versus the cylinder volume. This metric seems Elastomeric ring friction coefficient [74] 0.09 Elastomeric ring density [75] 2260 kg/m3 Elastomeric ring thermal conductivity [75] 0.560 W/m-K Elastomeric ring heat capacity [75] 1000 J/kg-K Table 2 – Ring-dynamic model parameters to provide a more rigorous means of evaluating the modeling effort compared to the traditional metric of comparing pressure traces with respect to time. Three maximum compression ratio cases are shown (25, 37 and 50:1). There are two sets of curves, the top set incorporates the ring-dynamic model into the calculations, while the bottom set does not include any crevice sub-model; the performance of the single-zone model is shown in a later plot. The curves are plotted from the experiment initialization through the first compression and expansion stroke, moving from right to left and back again on the graph. The bottom set of curves run along a counter-clockwise path through the experiment, as is indicated by the arrows. There is a slight shift in both sets of curves at the experiment startup. This is consistent throughout most of the RCEM data (with some shifts indicating experimental pressures higher than the ‘modified-entropy’ pressures), and could be due to some electrical noise in the system, but this is not yet understood. The bottom set of curves indicate the extent of the crevice flow in this experiment. It is apparent that as the charge is compressed to higher compression ratios more of the charge will be forced into the ringpack (up to 20%), thus the experiment pressure drops relative to the ‘modified-entropy’ pressure. The lag time for the return of the charge into the bulk volume is also visible. Not all of the mass reemerges from the ringpack by the end of Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 20 1.15 Initial shift in experimental pressure (due to noise?) 1.00 Air-only Charge Compression Ratio 50:1 Initial Temperature 25C Initial Pressure 0.099MPa 1.10 0.90 1.00 Drop in experimental pressure due to heat / mass transfer to ringpack 0.90 End of experiment a1_11, CR=25:1 a1_14, CR=37:1 a1_17, CR=50:1 0.70 0.65 0 200 400 0.85 Air-only Charge Compression Ratio 25,37,50:1 Initial Temperature 25C Initial Pressure 0.099MPa 600 800 1000 2 10 3 1 4 5 Elastomeric ring does not slip 0.80 3 Volume [cm ] Figure 9 – Ratio of experimental to modified-entropy pressure vs. cylinder volume illustrating the issue of crevice flows and the capability of the tuned ring-dynamic crevice model; ‘motored’ aironly shot, showing three maximum geometric compression ratios. the stroke, and there is substantial heat loss when the compressed gases come in contact with the crevice walls; these issues result in cumulative losses of about 10% for all of the runs. It is also clear that the pressure ratio is fairly stable through most of the compression stroke, indicating that the majority of the crevice flows don’t occur until after a compression ratio of about 5.5:1 (where the volume is 200cm3) is achieved. The top set of curves illustrates the capability of the ringdynamic sub-model to reasonably account for the crevice flows in the RCEM. The model parameters have been adjusted so that the pressure ratio is maintained at about 0.98 through both strokes of the piston. There is some fluctuation/noise in the trace, but the overall agreement is very good. One issue still unresolved however, is the slight drop in the pressure ratio on the expansion stroke for the higher CR runs; this was also apparent in Figs. 5a and 5b during the expansion from a compression ratio of 48:1 to 22:1. The drop seems to indicate that there is unpredicted ringpack charging / discharging on the expansion stroke. 5 0 0.75 1200 0 10 20 30 40 50 Instantaneous Compression Ratio Figure 10 – Ringpack pocket pressures vs. compression ratio for a ‘motored’ air-only shot, using a maximum geometric compression ratio of 50:1. 1000 Forces [kgf] 0.75 /P 0.80 calc 0.95 exp 0.85 Pressure [MPa] 1.05 Slight discrepancy for higher CR runs on expansion stroke, even with ring dynamic model. Cross-over of pressure leads to reemergence of charge, (from volume 3 to 1) 15 P P 0.95 Compression ring slippage 2500 'Fired' run, 29_06, Natural Gas / Air Equivalence Ratio 0.365 Initial Temperature 67C Initial Pressure 0.1MPa 500 2000 0 1500 -500 1000 -1000 500 -1500 0 -2000 -500 Friction force (F ) Forces [kgf] 1.05 exp /P calc Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM f1 'Motored' run, a1_17, Air-only Initial Temperature 25C Initial Pressure 0.099MPa Pressure force (F ) -2500 p1 Inertial force (Fi1) Net force (F +F +F ) -3000 f1 0 10 p1 -1000 i1 20 30 40 50 -1500 Instantaneous Compression Ratio Figure 11 – Forces on the compression ring vs. compression ratio for a ‘motored’ air-only and a ‘fired’ (natural gas / air) shot. Maximum compression ratios of 50:1 and 42:1, respectively, were used. 0.20 Air-only Charge Compression Ratio 25,37,50:1 Initial Temperature 25C Initial Pressure 0.099MPa Crevice Mass Fraction Figure 10 illustrates the behavior of the ringpack pocket pressures, for the model parameters used. This case is a high CR ‘motored’ run (the same as that shown in Figs. 5a & 5b) and the pressures are plotted versus instantaneous compression ratio; again arrows indicate the direction of the curves. The individual pocket pressures are labeled by the volume numbers. It is immediately clear that there is a lag in the pressurization/depressurization of the ringpack, as all of the curves are different from volume 1 which is at the bulk volume pressure. This is due to the narrowness of the transfer gaps available to the flow. Two points of interest are the drop in the volume 2 pressure as the compression ring shifts in the groove, and the cross-over point where the volume 3 pressure becomes greater than the pressure in volume 1, thus leading to the reemergence of the charge from the ringpack. The shift Rollover indicates point of reemergence from ringpack back into bulk charge 0.15 a1_17 Crossover at end of stroke for the 3 runs 0.10 a1_14 a1_11 0.05 0.00 0 10 20 30 40 50 Instantaneous Compression Ratio Figure 12 – Mass fraction of cylinder charge in the crevice and ringpack volumes vs. compression ratio for ‘motored’ air-only runs, using three different maximum geometric compression ratios. Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM Two other featuress are evident from these calculations. First, the reemergence of the crevice charge begins shortly after TDC has been reached. This is slightly earlier than the crossover of pressures between volumes 1 and 3, and is led by the transfer of the top-land charge back to the main volume. This early reemergence may be important if the chemical kinetics are slow and they occur on the expansion stroke; the reemerging cooler charge could significantly influence the bulk charge chemistry or the interpretation of it. Secondly, there is a cross-over in the mass fraction traces for the three runs presented. This cross-over is at the end of the runs and indicates that the operation with higher CR conditions results in a return of more charge to the cylinder on the expansion stroke of the experiment. In order to accomplish this with the 0-D model the circumferential area of the elastomeric ring needed to be adjusted over the range of CR conditions investigated (from 1.5mm2 to 1.0mm2); this was the only ringpack parameter that was adjusted. This issue seems to indicate that there is some inconsistency within the model that still needs to be resolved. One possibility which may explain this is if volume 5 is not at atmospheric conditions, but is pressurized during the run due to the presence of the rider ring. Higher CR runs will result in higher pocket Air-only Charge Compression Ratio 50:1 Initial Temperature 25C Initial Pressure 0.099MPa 1.10 Drop in computed pressure at end of experiment due to too much heat / mass transfer to ringpack /P calc 1.05 1.00 0.95 Initial shift in experimental pressure (due to noise?) exp Figure 11 illustrates the component forces (from Fig. 7) on the compression ring for both a ‘motored’ case and a ‘fired’ case. The inertial force is clearly the smallest of the forces for both cases and is negligible. In addition, it can be seen that the friction forces are much higher than those reported in previous studies (Refs. [51-53,59]). The behavior of the combined forces results in almost instantaneous slippage of the ring at TDC. This behavior is seen in both the ‘motored’ case and a ‘fired’ case presented here.Figure 12 shows the mass fraction of the cylinder charge that is driven into the top-land crevice and ringpack volumes for various ‘motored’ runs; three different maximum geometric compression ratio cases are presented. As was seen earlier in Fig. 9, less than 0.5% of the charge is forced into the crevice/ringpack before a compression ratio of 5.5:1 has been reached. As greater and greater compression is utilized (to achieve greater levels of compression heating), substantially more charge is driven into the ringpack. For the 50:1 run nearly 15% enters these volumes; this has a significant impact with regard to interpreting the cycle thermal efficiency, as a large fraction of the thermal energy is lost to the ringpack surfaces. A better option to achieve charge heating in future work might be to replace some of the nitrogen with a diluent like argon which has a higher specific heat ratio. 1.15 P in the compression ring position is interesting here since it occurs at nearly TDC and very quickly. This is due to the high friction forces on the ring (as can be seen in Fig. 11) which result from the unlubricated conditions, and the high friction coefficient for this ring (Cf1 = 0.24). This is different from the results reported in previous studies (e.g., [52,53,57]). End of experiment 0.90 0.85 ring dynamic model, with ring heating no crevice model ring dynamic model, no ring heating, circum. area (A35c ) =0 0.80 0.75 ring dynamic model, no ring heating, circum. area (A35c ) =1.1mm single-zone model 0 200 400 600 800 1000 2 1200 3 Volume [cm ] Figure 13 – Ratio of experimental to computed pressures vs. cylinder volume illustrating the effects of various modeling parameters for a ‘motored’ air-only shot, using a maximum geometric compression ratio of 50:1. pressures and more charge will be returned to the cylinder. Figure 13 is presented to illustrate some of the effects that various modeling parameters have on the computed ‘modified-entropy’ pressure. Here the pressure ratio is shown again versus cylinder volume. A maximum geometric compression ratio of 50:1 is used for this example. Traces are presented for the tuned ringdynamic model (parameters listed in Table 2), the case with no crevice sub-modeling (for reference), a calculation using the single-zone crevice sub-model, and two cases that do not include ring heating (for the ringdynamic model). It can be clearly seen that the singlezone model cannot accurately replicate the dynamics of the crevice flows (as was evident earlier in Figs. 5a & 5b). In addition, it evident that although the ringdynamic model without ring heating can accurately predict the cylinder pressure on the compression stroke, this model option results in a computed ‘modifiedentropy’ pressure that is lower than the experimentally measure pressure on the expansion stroke. Modifying the circumferential area, A35c to 0.0 seems to improve the match but there is still a discrepancy. These results indicate that in order to match the compression pressure, a certain fraction of mass and energy must be removed from the cylinder. However if all of the sensible enthalpy of this mass is transferred to the walls then the computed pressure will be too low. Thus, it is thought that ring heating is an important parameter that can account for this. To illustrate the extent of the friction heating, examples of selected temperature profiles within the compression and elastomeric rings are presented in Figures 14a & 14b; the computed temperature profiles are plotted versus time for different locations within the grid that are close to the ring surface. ‘Motored’ conditions have been used for illustration at two different compression Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM 1100 Air-only Charge Compression Ratio 25:1 Initial Temperature 25C Initial Pressure 0.099MPa 800 1000 700 900 Ring Surface 10 Pm from surface 50 Pm from surface 100 Pm from surface 600 500 Compression Ring TDC 800 700 Into ring 400 600 300 500 Into ring 200 100 400 Elastomeric Ring 0 5 10 15 20 25 30 300 Temperature [K] Temperature [K] 900 Time [ms] Figure 14a – Temperature profiles within the rings of a ‘motored’ air-only shot, using a maximum geometric compression ratio of 25:1. 1100 Air-only Charge Compression Ratio 50:1 Initial Temperature 25C Initial Pressure 0.099MPa 800 Compression Ring 1000 700 900 Ring Surface 10 Pm from surface 50 Pm from surface 100 Pm from surface 600 Into ring 800 500 700 TDC 400 300 500 Into ring Elastomeric Ring 200 100 600 0 5 10 400 15 20 25 30 300 Temperature [K] Temperature [K] 900 pressures are still high on the expansion stroke (as seen in Fig. 10) and this results in high friction forces. Uncertainties exist in the calculated ring temperatures; these may be associated with the evaluation of the ring thermal diffusivity (there could be substantial changes as the temperature increases) and the fact that as the ring surface becomes significantly hot, a higher fraction of the heat will be transferred away from the colder ring to the cylinder wall and not to the ring’s interior (the constant value of the friction fraction may not be realistic). ‘FIRED’ RUNS The integrated ring-dynamic crevice model and multizonal heat release code has been demonstrated to provide a reasonably good match between the computed ‘modified-entropy’ and experimental cylinder pressures for a number of ‘motored’ cases. The purpose of this section is to apply this integrated code to two ‘fired’ cases to illustrate how the crevice dynamics change under the ‘fired’ conditions, and how these overall dynamics can influence the accuracy and interpretation of the autoignition/HCCI data. Figures 15a & 15b present the computed ringpack pocket pressures for two ‘fired’ natural gas / air runs where the ignition occurs before TDC (early) and after TDC (late), respectively. These plots are schematically similar to Fig. 10. One notable feature is that the model predicts that a significant amount of mass can be driven into crevice and ringpack as the autoignition process occurs (this is observable from the volume 1 and volume 3 pressure traces). The pressure in volume 3 more than doubles in Fig. 15a as the main charge autoignites. Time [ms] Figure 14b – Temperature profiles within the rings of a ‘motored’ air-only shot, using a maximum geometric compression ratio of 50:1. ratios; higher compression ratios will result in higher ringpack pocket pressures and thus higher friction forces on the rings. The heating will be more significant for high CR and boosted cases, as well as for ‘fired’ runs. The computed temperature rise at the ring surface is substantial due to the low thermal diffusivity of the rings used in these experiments (Įcomp ring = 0.35x10-6m2/s); there is little (slow) penetration of the heat into the ring, leaving it concentrated at the ring surface. Conventional metal rings are more effective at diffusing the friction heat away from the ring surface (Įsteel = 15x10-6m2/s). On a typical lubricated surface a maximum temperature rise of less than 20K for typical compression rings would be predicted for the pressure profiles in the RCEM; less than 60K would be seen on an unlubricated surface. In Figs. 14a and 14b it can be seen that there is a slight drop in the ring surface temperatures near TDC as the piston comes to rest, and then there is additional heating as the piston reverses direction. The ringpack pocket Another interesting aspect can be interpreted from Figs. 15 and 16, which shows the crevice mass fraction for the two ‘fired’ runs versus instantaneous geometric compression ratio. It appears that for both cases, whether the charge ignites before TDC or after TDC, the crevice mass does not reemerge into the main charge until after the autoignition event is essentially complete (for these natural gas / air runs). This point, if real, is beneficial with regard to accurately characterizing the bulk charge temperature, especially for late-firing conditions (there would not be a significant mass of cold reemerging crevice charge during the ignition process; more investigation is needed to see if similar behavior is seen for very-late ignition or slower reacting fuel / air charges. Also seen in Fig. 16 is the crossover issue that was discussed for the ‘motored’ runs. Again, the circumferential gap area (A35c) of the elastomeric ring was adjusted so that agreement could be achieved for the pressure ratio (Pexp / Pcalc) late in the expansion stroke; as with the ‘motored’ cases, this was the only ringpack parameter adjusted between these runs. Figure 17 is presented next to illustrate the importance of accurately characterizing the crevice flow behavior in Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM Compression ring slippage Tempertature [K] Cross-over of pressure leads to reemergence of charge, (from volume 3 to 1) 15 Ignition Elastomeric ring does not slip 10 1 2 5 1500 2800 Discrepancy through expansion 1000 Discrepancy at ignition 500 10 20 30 -1000 40 Instantaneous Compression Ratio 29_11 Compression ring slippage Cross-over of pressure leads to reemergence of charge, (from volume 3 to 1) Ignition 2 1 3 4 10 20 30 Instantaneous Compression Ratio Figure 15b – Ringpack pocket pressures vs. compression ratio for a ‘fired’ natural gas / air shot, using a maximum geometric compression 32:1. 0.20 Natural Gas / Air Equivalence Ratio 0.365 Initial Temperature 67C Initial Pressure 0.1MPa 29_06 29_11 0.10 25 30 35 40 0.1 0.0 29_11 -0.3 -0.1 ignition after TDC -0.4 0.0 -0.5 -0.1 -0.6 -0.2 -0.7 -0.2 29_06 -0.3 Greater crevice loss due to more crevice flow ignition before TDC -0.4 heat loss to walls (modified Woschni) heat loss in crevice combined heat loss (wall + crevice) 0 5 10 15 -0.5 20 25 Time [ms] Figure 18 – Calculated heat loss from the bulk charge for a ‘motored’ run and two ‘fired’ runs. The issue of ring heating is observable. crevice flow accounts for difference in RCEM KTH and ideal calculations close to emissions data thermodynamic cycle 1.00 2.00 0.75 1.75 0.50 29_11 ring-dynamic 0.25 29_06 ring-dynamic 0.00 behavior matches crevice flow dynamics 1.50 1.25 discrepancy with emissions data burned fraction in cylinder burned fraction overall 1.00 0.75 -0.50 -0.75 0.05 0.50 29_11 single-zone -1.00 29_06 single-zone negative burned fraction behavior does not match single-zone crevice flows, or current HCCI understanding 0.25 0.00 -1.25 -2 0.00 0 10 20 30 300 0.2 Positive heat transfer from heated ring -0.25 Rollover indicates point of reemergence from ringpack back into bulk charge Crossover at end of stroke 20 -0.2 40 0.15 15 -0.1 -0.8 0 10 a1_17 Elastomeric ring does not slip 5 0 Cumulative Heat Loss [kJ] 15 Mass Fraction Burned Pressure [MPa] 29_11 5 5 0.0 Natural Gas / Air Equivalence Ratio 0.365 Initial Temperature 67C Initial Pressure 0.1MPa 10 0 Figure 17 – Mass-average temperature versus effective compression ratio for two ‘fired’ natural gas / air charges, illustrating the effects of using the ring-dynamic and single-zone models. 25 Crevice Mass Fraction 800 Average temperature, ring dynamic model Average temperature, single zone model Instantaneous Effective Compression Ratio Figure 15a – Ringpack pocket pressures vs. compression ratio for a ‘fired’ natural gas / air shot, using a maximum geometric compression 42:1. 20 1800 1300 -500 0 2300 29_06 5 0 3300 0 3 4 2000 40 Instantaneous Compression Ratio Figure 16 – Mass fraction of cylinder charge in the crevice and ringpack volumes versus compression ratio for ‘fired’ natural gas / air charges, using two different maximum compression ratios. Cumulative Heat Loss [kJ] 29_06 3800 Natural Gas / Air Equivalence Ratio 0.365 Initial Temperature 67C Initial Pressure 0.1MPa 0 2 4 6 0 8 2 10 4 12 6 14 Mass Fraction Burned 20 Pressure [MPa] 2500 Natural Gas / Air Equivalence Ratio 0.365 Initial Temperature 67C Initial Pressure 0.1MPa Temperature [K] 25 Time (arbitrary) [ms] Figure 19 – Calculated mass fraction burned versus time for two ‘fired’ natural gas / air charges, illustrating the effects of using the ring-dynamic and single-zone crevice models. Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM the RCEM. The mass-average temperatures of two ‘fired’ runs are shown versus instantaneous effective compression ratio; the effective compression ratio is the ratio of the mass-specific volumes, as opposed to the ratio of the geometric volumes ( CR eff X ini X inst ). This takes into account the mass transfer to and from the bulk charge (from & to the crevice/ringpack volume). The temperature curves have been modified slightly to correct for the shift in the pressure data that occurs at the experiment initiation, which was discussed in reference to Fig. 9. Calculations using the tuned ringdynamic model and the single-zone model are presented in Fig. 17. From this plot it is apparent that both the autoignition temperature and the temperature on the expansion stroke are significantly impacted by the crevice modeling. The accuracy of the autoignition temperature is extremely important with regard to kinetic modeling. The expansion temperature is related to the mass fraction burned calculations; lower computed temperatures will indicate less charge oxidation, and this will lead to discrepancies between the calculations and any emissions measurements that are taken. Figure 18 plots the cumulative heat loss for three runs, one ‘motored’ and two ‘fired’ cases. The bulk charge (modified Woschni correlation) and the crevice heat losses are shown independently, the combined heat loss is also plotted. It is apparent that the crevice losses are significantly greater than the bulk charge losses; the crevice losses are on the order of 15% of the fuel energy while the main cylinder losses are only about 3% for the fired cases (~1% for the ‘motored’ case). The effect of the ring heating modeling is also visible in the ‘motored’ case, as there is apparently a decrease in the magnitude of the crevice heat loss as the crevice charge reemerges into the main cylinder. Finally, it can be seen that the case where ignition occurs before TDC seems to have greater heat loss in the crevice due to the increase in crevice flow (as was seen in Fig. 16). An increased loss of about 4% of the fuel energy is observed. However, this result does not necessarily match the thermal efficiency trends presented in Ref. [7] where there seemed to be little variation in the computed thermal efficiency with increased compression ratio. More investigation of this point is required. Finally, Figure 19 presents the results of the mass fractions burned calculations for the two ‘fired’ natural gas / air cases. Comparison is made between the ringdynamic model and the single-zone models. Two curves are shown for each case, one is an overall value of the mass fraction burned, and the other for the fraction of the charge that is within the bulk volume that is determined to be oxidized. Again, this is calculated by matching the ‘modified-entropy’ pressure to the experimental pressure; the heat released from combustion is required to account for the pressure change not associated with piston compression & expansion, heat loss to the walls, or mass transfer to/from the crevice. The discrepancies of the single-zone approach are again evident. First, there is a computed negative burned fraction just before the main heat release; this could be interpreted incorrectly as an endothermic preignition period. Secondly, there appears to be a spike and then a drop in the mass fraction burned, especially for the over-compressed case (29_06); the charge then appears to oxidize more as the time progresses. This might be interpreted (again incorrectly) as a high pressure effect, since the cylinder pressure is significantly greater for this run. Finally, the computed burned fractions indicate that only 80% of the charge has oxidized, while the emissions measurements for these runs indicate that close to 98% of the of the fuel has burned. This discrepancy may be due to blowby mass however. The ring-dynamic model seems to provide more plausible results compared with the single zone model, and the effects of the crevice flows can be clearly seen (especially for the 29_06 case). The computed mass fractions burned (for the main cylinder charge) much more closely match the emissions data, and the overall values seem to account for the amount of mass that is calculated to be trapped in the ringpack at the end of the stroke, as well as the heat that is lost to the crevice/ringpack walls. The difference between the thermodynamic cycle efficiency of the RCEM experiments, 56% as presented in Ref. [7], and the ideal Otto cycle efficiency for these compression ratios (up to 68%) is more understandable with the use of the ringdynamic model. SUMMARY AND CONCLUSIONS A ring-dynamic crevice model has been modified and integrated with a multi-zonal heat release code to investigate the effects of crevice flows on the interpretation of autoignition (HCCI) data from a Rapid Compression Expansion Machine. The model parameters have been tuned so that reasonable agreement could be achieved for a number of ‘motored’, air-only compression – expansion runs. The ratio of the experimentally measured pressure to the computed, ‘modified-entropy’ pressure provides a rigorous metric for evaluating the performance of the crevice sub-model. It was found that the heating of the low thermal diffusivity rings due to sliding friction as the piston traverses the unlubricated walls of the cylinder, may significantly influence the mass flow rates and the enthalpy of the reemerging crevice flows. A significant fraction of the cylinder charge is computed to be driven into the crevice/ringpack (15%) (since there is no wet seal) as high compression ratios (50:1) are achieved. The modified ring-dynamic model has been able to provide useful insight with regard to the interpreting ‘fired’ runs as well. The characteristics of the autoignition (HCCI) process, including the autoignition temperature, the rate of heat release and the computed mass fraction burned are much better understood and the experimental results seem to better correlate with Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM current HCCI understanding. Additional studies, including the development and refinement of various kinetic models will now be possible. An improved ringpack for the RCEM, which could minimize crevice flows and/or prevent the cooler crevice charge from reentering the bulk volume on the expansion stroke, might improve the capabilities of this experimental apparatus. This study has also provided some understanding with regard to the discrepancies between the ideal Otto cycle performance, and the efficiencies that have been computed for this experiment. If the crevice flows can be controlled, along with the associated heat losses within the crevice and ringpack, through sufficient sealing, it appears that even better cycle performance could be achieved. ACKNOWLEDGMENTS The author is grateful to Peter Van Blarigan and Nick Paradiso of Sandia National Laboratories, California who provided the raw data and experimental details used for this work. REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. Park, P. and Keck. J.C., “Ignition Delays for Iso-octane: Measurement Using a Rapid Compression Machine and Prediction Using a Reduced Chemical Kinetic Model,” SAE Paper 912553, 1991. Griffiths, J.F., Halford-Maw, P.A. and Rose, D.J., “Fundamental Features of Hydrocarbon Autoignition in a Rapid Compression Machine,” Combustion and Flame, 95:3, 291-306, 1993. Furutani, M., Ohta, Y. and Komatsu, K., “Onset Behavior of Low-Temperature Flames Caused by Piston Compression,” JSAE Review, 14:12-17, 1993. Minetti, R., Carlier, M., Ribaucour, M., Therssen, E. and Sochet, L.R., “A Rapid Compression Machine Investigation of Oxidation and Auto-ignition of n-Heptane: Measurements and Modeling,” Combustion and Flame, 102:298-209, 1995. Murase, E., Ono, S., Hanada, K., Yun, J.H., Oppenheim, A.K., “Performance of pulsed combustion jet at high pressures and temperature,” JSAE Review, 17:245-250, 1996. Griffiths, J.F., Halford-Maw, P.A. and Mohamed, C., “Spontaneous Ignition Delays as a Diagnostic of the Propensity of Alkanes to Cause Engine Knock,” Combustion and Flame 111:327-337, 1997. Van Blarigan, P., Paradiso, N. and Goldsborough, S., “Homogeneous Charge Compression Ignition with a Free Piston: A New Approach to Ideal Otto Cycle Performance,” SAE Paper 982484, 1998. Minetti, R., Roubaud, A., Therssen, E., Ribaucour, M. and Sochet, L.R., “The chemistry of pre-ignition of n-pentane and 1-pentene,” Combustion and Flame, 118:213-220, 1999. 9. 10. 11. 12. 13. Roubaud, A. Minetti, R. and Sochet, L.R., “Oxidation and combustion of low alkylbenzenes at high pressure: comparative reactivity and auto-ignition,” Combustion and Flame 121:535-541, 2000. Roubaud, A., Lemaire, O., Minetti, R. and Sochet, L.R., “High pressure auto-ignition and oxidation mechanisms of o-xylene, o-ethyltoluene, and n-butylbenzene between 600 and 900 K,” Combustion and Flame, 123:561-571, 2000. Lemaire, O., Ribaucour, M., Carlier, M. and Minetti R., “The production of benzene in the low-temperature oxidation of cyclohexane, cyclohexene, and cyclohexa-1,3diene,” Combustion and Flame 127:1971-1980, 2001. Griffiths, J.F. and Whitaker, B.J., “Thermokinetic Interactions Leading to Knock During Homogenous Charge Compression Ignition,” Combustion and Flame 131:386-399, 2002. Griffiths, J.F., MacNamara, J.P., Sheppard, C.G.W., Turton, D.A. and Whitaker, B.J., “The relationship of knock during controlled autoignition to temperature inhomogeneities and fuel reactivity,” Fuel 81:2219-2225, 2002. 14. Tanaka, S., Ayala, F., Keck, J.C. and Heywood, J.B., “Two-stage ignition in HCCI combustion and HCCI control by fuels and additives,” Combustion and Flame 132:219239, 2003. 15. Ando, T., Isobe, Y., Sunohara, D., Daisho, Y. and Kusaka, J., “Homogeneous charge compression ignition and combustion characteristics of natural gas mixtures: the visualization and analysis of combustion,” JSAE Review, 24:33-40, 2003. 16. Lim, O.T., Sendoh, N. and Iida, N., “Experimental Study on HCCI Combustion Characteristics of n-Heptane and isoOctane Fuel/Air Mixture by the use of a Rapid Compression Machine,” SAE Paper 2004-01-1968, 2004. 17. Silke, E. J., Curran H. J. and Simmie J. M., “The Influence of Fuel Structure on Combustion as Demonstrated by the Isomers of Heptane; a Rapid Compression Machine Study,” Proceedings of 30th International Symposium on Combustion, Chicago, July 2004. 18. Donovan, M. T., He X., Zigler, B. T., Palmer, T. R., Wooldridge, M. S. and Atreya A., “Demonstration of a Free-Piston Rapid Compression Facility for the Study of High Temperature Combustion Phenomena,” Combustion and Flame 137:351-365, 2004. 19. He, X., Donovan, M.T., Zigler, B.T., Palmer, T.R., Walton, S.M., Wooldridge, M.S. and Atreya, A., “An experimental and modeling study of iso-octane ignition delay times under homogeneous charge compression ignition conditions,” Combustion and Flame, 142:266-275, 2005. 20. Murase, E., Hanada, K., Miyaura, T., and Ikeda, J., “Photographic Observation and Emission Spectral Analysis of Homogeneous Charge Compression Ignition Combustion,” Combustion Science and Technology, 177:1699-1723, 2005. 21. Vanhove, G., Minetti, R., Touchard, S., Fournet, R., Glaude, P.A. and Battin-Leclerc, F., “Experimental and modeling study of the autoignition of 1-hexene/isooctane mixtures at low temperatures,” Combustion and Flame, 145:272-281, 2006. Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM 22. He, X., Zigler, B.T., Walton, S.M., Wooldridge, M.S., and Atreya, A., “A rapid compression facility study of OH time histories during iso-octane ignition,” Combustion and Flame, 145:552-570, 2006. 23. Vanhove, G., Petit, G., Minetti, R., ”Experimental study of the kinetic interatctions in the low-temperature autoignition of hydrocarbon binary mixtures and a surrogate fuel,” Combustion and Flame, 145:521-532, 2006. 24. Mittal, G., Sung, C.J. and Yetter, R.A., “Autoignition of H2/CO at elevated pressures in a rapid compression machine,” International Journal of Chemical Kinetics, 38:516-529, 2006. 25. Griffiths, J.F., Hughes, K.J., Schreiber, M. and Poppe, C., “A Unified Approach to the Reduced Kinetic Modeling of Alkane Combustion,” Combustion and Flame, 99:533-540, 1994. 26. Faravelli, T., Gaffuri, P., Ranzi, E. and Griffiths, J.F., “Detailed thermokinetic modeling of alkane autoignition as a tool for the optimization of performance of internal combustion engines,” Fuel 77:147-155, 1998. 27. Minetti, R., Roubaud, A., Therssen, E., Ribaucour, M. and Sochet, L.R., “The chemistry of pre-ignition of n-pentane and 1-pentene,” Combustion and Flame, 118:213-220, 1999. 28. Ribaucour, M., Minetti, R., Sochet, L.R., Curran, H., Pitz, W.J., and Westbrook, C.K., “Ignition of isomers of pentane; an experimental and modeling study, Proceedings of the Combustion Institute, 28:1671-1678, 2000. 29. Westbrook, C.K., Pitz, W.J., Boercker, J.E., Curran, H.J., Griffiths, J.F., Mohamed, C. and Ribaucour, M., “Detailed Chemical Kinetic Reaction Mechanisms for Autoignition of Isomers of Heptane under Rapid Compression,” Proceedings of the Combustion Institute, 29:1311-1318, 2002. 30. Curran, H. J., Gaffuri, P., Pitz, W. J., and Westbrook, C. K. "A Comprehensive Modeling Study of iso-Octane Oxidation" Combustion and Flame 129:253-280, 2002. 31. Tanaka, S., Ayala, F. and Keck, J.C., “A reduced chemical kinetic model for HCCI combustion of primary reference fuels in a rapid compression machine,” Combustion and Flame 133:467-481, 2003. 32. Curran, H. J., Gaffuri, P., Pitz, W. J., and Westbrook, C. K. "A Comprehensive Modeling Study of n-Heptane Oxidation" Combustion and Flame 114:149-177, 1998. 33. Buda, F., Bounaceur, R., Warth, V., Glaude, P.A., Rournet, R. and Battin-Leclerc, F., “Progress toward a unified detailed kinetic model for the autoignition of alkanes from C4 to C10 between 600 and 1200K,” Combustion and Flame, 142:170-186, 2005. 34. Potokar, C. and Goldsborough, S.S., “The influence of crevice flows and blow-by on the charge motion and temperature profiles within a Rapid Compression Expansion Machine used for chemical kinetic (HCCI) studies,” SAE Paper 2007-01-0169, 2007. 35. Goldsborough, S.S., “Evaluating the Heat Losses from HCCI Combustion within a Rapid Compression Expansion Machine,” SAE Paper 2006-01-0870, 2006. 36. Griffiths, J.F, MacNamara, J.P., Mohamed, C., Whitaker, B.J., Pan J. and Sheppard, C.G.W., “Temperature fields 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. during the development of autoignition in a rapid compression machine,” Faraday Discuss., 119:287-303, 2001. Clarkson, J., Griffiths, J.F., Macnamara, J.P. and Whitaker, B.J., “Temperature Fields During the Development of Combustion in a Rapid Compression Machine,” Combustion and Flame 125:1162-1175, 2001. Mittal, G. and Sung, C.J., “Aerodynamics inside a rapid compression machine,” Combustion and Flame, 145:160180, 2006. Brett, L., Macnamara, J., Musch, P. and Simmie, J.M., “Simulation of Methane Autoignition in a Rapid Compression Machine with Creviced Pistons,” Combustion and Flame 124:326-329, 2001. Meere, M.G., Gleeson, B. and Simmie, J.M., “Modelling gas motion in a rapid-compression machine,” Journal of Engineering Mathematics 44:57-82, 2002. Würmel, J. and Simmie, J.M., “CFD studies of a twinpiston rapid compression machine,” Combustion and Flame, 141:417-430, 2005. Lee, D. and Hochgreb, S., “Rapid Compression Machines: Heat Transfer and Suppression of Corner Vortex,” Combustion and Flame 114:531-545, 1998. Tabaczynski, R.J., Hoult, D.P. and Keck, J.C., “High Reynolds Number Flow in a Moving Corner,” Journal of Fluid Mechanics, 42:249-255, 1970. Daneshyar, H.F., Fuller, D.E. and Deckker, B.E.L., “Vortex Motion Induced by the Piston of an Internal Combustion Engine,” International Journal of Mechanical Sciences, 15:381-390, 1973. Park, J.W., Huh, K.Y. and Park, K.H., “Experimental study on the combustion characteristics of emulsified diesel in a rapid compression and expansion machine,” Proceedings of the Institution of Mechanical Engineers 214, Part D, 2000. Silke, E., PhD Dissertation, Chemistry Department, National University of Ireland – Galway, 2005. Huang, Z., Shiga, S., Ueda, T., Nakamura, H., Ishima, T., Obokata, T., Tsue, M. and Kono, M., “Effect of Fuel Injection Timing Relative to Ignition Timing on the NaturalGas Direct-Injection Combustion,” Journal of Engineering for Gas Turbines and Power, 125:783-790, 2003. Furuhama, S. and Tada, T., “On the Flow of Gas Through the Piston-Rings. 1st Report, The Discharge Coefficient and Temperature of Leakage Gas,” Bull JSME, 4:16, 1961. Chang, J., Güralp, O., Filipi, Z., Assanis, D., Kuo, T.W., Najt P. and Rask, R., “New Heat Transfer Correlation for an HCCI Engine Derived from measurements of Instantaneous Surface Heat Flux,” SAE Paper 2004-012996, 2004. Chang, J., Filipi, Z., Assanis, D. Kuo, T-W, Najt, P., Rask, R., “Characterizing the thermal sensitivity of a gasoline homogeneous charge compression ignition engine with measurements of instantaneous wall temperature and heat flux,” International Journal of Engine Research 6:289-309, 2005. Namazian, M., “Studies of Combustion and Crevice Gas Motion in a Flow-Visualization Spark-Ignition Engine, PhD Dissertation, Massachusetts Institute of Technology, 1981. Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM 52. Namazian, M. and Heywood, J.B., “Flow in the PistonCylinder-Ring Crevices of a Spark-Ignition Engine: Effect on Hydrocarbon Emissions, Efficiency and Power,” SAE Paper 820088, 1982. 53. Kuo, T.W., Sellnau, M.C., Theobald, M.A. and Jones, J.D., “Calculation of Flow in the Piston-Cylinder-Ring Crevices of a Homogenous-Charge Engine and Comparison with Experiment,” SAE Paper 890838, 1989. 54. Reitz, R. and Kuo, T.W., “Modeling of HC Emissions Due to Crevice Flows in Premixed-Charge Engines,” SAE Paper 892085, 1989. 55. Amsden, A.A, Ramshaw, J.D., O-Rourke, P.J. and Dukowicz, J.K., “KIVA: A Computer Program for Two-And Three-Dimensional Fluid Flows with Chemical Reactions and Fuel Sprays,” Los Alamos Scientific Laboratory Report LA-10245-MS, February 1985. 56. Tonse, S.R., “Numerical Simulations of Emerging Piston Crevice Gases,” SAE Paper 961968, 1996. 57. Huynh, MM.C., Gatellier, B., Baritaud, T., Dagand, M.F., and Galliot, M.F., “Modelisation 3D de la Formation des Hydrocarbures Imbrules dan les Moteurs a Allumage Commande,” SAE Paper 963409, 1996. 58. Roberts, C.E. and Matthews, R.D., “Development and Application of an Improved Ring Pack Model for Hydrocarbon Emissions Studies,” SAE Paper 961966, 1996. 59. Zhao, J.X. and Lee, C.F., “Modeling of Blow-by in a SmallBore High-Speed Direct-Injection Optically Accessible Diesel Engine,” SAE Paper 2006-01-0649, 2006. 60. Mansouri, S.H. and Heywood, J.B., “Correlations for the Viscosity and Prandtl Number of Hydrocarbon-Air Combustion Products,” Combustion Science and Technology, 23:251-256, 1980. 61. Krieger, R.B. and Borman, G.L., “The Computation of Apparent Heat Release for Internal Combustion Engines,” ASME Paper 66-WA/DGP-4, 1966. 62. Gatowski, J.A., Balles, E.N., Chun, K.M., Nelson, F.E., Ekchian, J.A. and Heywood, J.B., “Heat Release Analysis of Engine Pressure Data,” SAE Paper 841359, 1984. 63. Heywood, J.B., Internal Combustion Engine Fundamentals, McGraw-Hill, 1988. 64. Heywood, J.B. and Chun, K.M., “Estimating Heat-Release and Mass-of Mixture Burned from Spark-Ignition Engine Pressure Data,” Combustion Science and Technology, 54:133-143, 1987. 65. Jensen, T.K. and Schrammof, J., “Three-zone Heat Release Model for Combustion Analysis in a Natural Gas SI Engine: Effects of Crevices and Cyclic Variations on UHC Emissions,” SAE Paper 2000-01-2802, 2000. 66. Grimm, B.M., Johnson, R.T., “Review of Simple Heat Release Computations,” SAE Paper 900445, 1990. 67. Homsy, S.C., Atreya, A., “An Experimental Heat Release Rate Analysis of a Diesel Engine Operating under Steady State Conditions,” SAE Paper 970889, 1997. 68. Brunt, M.F., Rai, H., Emtage, A.L., “The Calculation of Heat Release Energy from Engine Cylinder Pressure Data,” SAE Paper 981052, 1998. 69. Ball, J.K, Raine, R.R. and Stone, C.R., “Combustion analysis and cycle-by-cycle variations in spark ignition engine combustion. Part I: an evaluation of combustion 70. 71. 72. 73. 74. analysis routines by reference to model data,” Proceedings of the Institution of Mechanical Engineers 212, Part D, 1998. Guezennec, Y.G. and Hamama, W., “Two-zone Heat Release Analysis of Combustion Data and Calibration of Heat Transfer Correlation in an I.C. Engine,” SAE Paper 1999-01-0218, 1999. Woschni, G., “A Universally Applicable Equations for the Instantaneous Heat Transfer Coefficient in the Internal Combustion Engine,” SAE Paper 670931, 1967. “Material Data: Polyimide – High Performance Polymer, 05-127,” Cook. MatWeb, The Online Materials Database, Polyimide, Graphite Filled “Fluoroloy A27,” Sait-Gobain Performance Plastics, http://www.seals.saint-gobain.com/Media/Documents/ S0000000000000001018/FLOY%20A27.pdf CONTACT S. Scott Goldsborough Department of Mechanical Engineering Marquette University Milwaukee, Wisconsin 53201-1881 Scott.Goldsborough@mu.edu DEFINITIONS, ACRONYMS, ABBREVIATIONS Roman letters A , Area (m2) a c , acceleration in inertial reference frame (m/s2) BDC , bottom dead center c , speed of sound Cd , discharge coefficient Cf , friction coefficient CP , constant for viscosity fit CR , compression ratio ct , clearance height (m) d , diameter of ring (m) Ff , friction force (N) Fi , inertia force (N) Foil , oil film force (N) FP , pressure force (N) f fric , friction power factor H , cylinder characteristic dimension (m) h , specific enthalpy (kJ/kg) k , thermal conductivity (W/m-K) L , length (m) m , mass (g) , mass flow rate (g/s) m P , pressure (kPa) q w , convective heat flux (W/m2) R , universal gas constant (kJ/kg-K) S p , mean piston speed (m/s) s , specific entropy (kJ/kg-K) T , temperature (K) Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM t , time (s) TDC , top dead center 't , time step (s) v , velocity (m/s) V , volume (m3) , rate of work transfer (kJ/s) W x , spatial dimension (m) z c , axial position in inertial reference frame (m) Greek letters D , thermal diffusivity (m2/s) J , ratio of specific heats U , density (kg/m3) P , dynamic viscosity (N-s/m2) X , specific volume (m3/kg) Subscripts 0 , beginning of experiment 1 , crevice zonal index 1 2 , from zone 1 to zone 2 br , burned zone c , combustion calc , calculated, “modified entropy method” conv , convective cr , crevice zone d , downstream eff , effective ex , exiting exp , experiment fr , fresh zone g , gas in , incoming ini , conditions at start of experiment inst , instantaneous conditions motor , motored mx , mixed zone o , stagnation conditions r , ring ref , reference condition s , isentropic t , previous time step t + 't , new time u , upstream w , wall APPENDIX This section discusses briefly the fitting of the ringdynamic model and the sensitivity of the adjusted parameters. The model consists of a number of physical dimensions and an empirical constant (the friction power factor), all of which were described earlier. The values used in the tuned model are listed in Table 2 of the main text. The parameters tuned for this study were the ringpack volumes, the ringside clearance heights and the circumferential areas. The endgap areas were set to 0.0 since there is no gap for the elastomeric ring (this is a continuous piece) and the circumferential flows seem to dominate in this configuration (much better fitting of the model to the data is possible using this assumption). In the model the ringpack volumes determine the amount of charge that can leave the main cylinder; larger volumes are able to capture more mass. The flow areas determine the rate of ringpack pressurization. Insufficient areas restrict the flow and do not allow sufficient charging of the pockets; areas too large result in pressure equilibrium with the bulk gas – there is no lag in the pocket pressurization. The friction power mainly affects the reemerging charge temperature; the computed flow rates are also affected, but these can be modified by slightly adjusting the circumferential areas. In tuning the model the objective was to match both the compression and expansion stroke pressures; the former is critical to characterizing the ignition conditions for early firing, the latter impacts late cycle ignition / oxidation and overall heat release from the charge. The performance of the model was assessed by analyzing the ratio of the experimental to calculated pressures (Pexp / Pcalc) with respect to time, volume and instantaneous compression ratio; it is critical that the conditions near TDC are well matched. A fitness parameter was defined in order to quantify the capability of the model; this is expressed in Equation A1. The absolute value of the relative difference between the experimental and calculated values is integrated over the compression and expansion strokes. A low value of the fitness is desired. fitness ³ 0.75 CR exp ³ 0.25 1.0 ³ CR exp 1.0 V exp Vini ³ Pexp 1 dCR Pcalc dCR Pexp Pcalc V exp Vini 1 dV (A1) dV Figure A1 plots this fitness parameter with respect to time for a typical ‘motored’ shot; the 50:1 CR case discussed earlier is used as the example. This provides a rigorous test of the model since higher CR runs result in more crevice flows and are therefore more difficult to match. Three features are evident in this plot. First, the fitness is poor during the experiment initiation; this is due to the low signal-to-noise ratio at low pressure. Next, it can be seen that the fitness improves along the compression stroke to TDC indicating good agreement during the critical part of the experiment. The fitness then decreases just after TDC (significantly for this high CR case) due to the departure of the experimental trace from the calculated trace (as seen in Fig. 5); this shift in fitness is not as significant for the low CR shots. The sensitivities of the adjustable model parameters are presented in Figure A2. These are graphed as a percent where this is defined by Equation A2. The percent sensitivity is the relative difference in fitness with Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192 Licensed to Marquette University Licensed from the SAE Digital Library Copyright 2010 SAE International E-mailing, copying and internet posting are prohibited Downloaded Tuesday, January 26, 2010 5:02:45 PM 0.030 Air-only Charge Compression Ratio 50:1 Initial Temperature 25C Initial Pressure 0.099MPa 0.025 Low signal-to-noise ratio at experiment initiation Fitness 0.020 0.015 0.005 0.000 1 vol2 2 vol3 3 vol4 4 ct1 5 A 6 ct2 7 A 8 f 9 13c Departure from experimental pressure on expansion for high CR cases 0.010 vol1 35c TDC -5 0 5 10 15 20 25 30 Time [ms] Figure A1 – Fitness of ring-dynamic model vs. time based on fitness definition. the base fitness compared to the new fitness; the base value is used in the numerator to give a positive sensitivity for better fits and a negative sensitivity for worse fits. Each of the parameters is individually changed by 15 percent (increased and decreased) about the base value and the fitness of the model is recalculated. Values at TDC and at the end of the experiment are graphed. sensitivity ª fitness base º 1» 100 « «¬ fitness new »¼ (A2) fric Air-only Charge Compression Ratio 50:1 Initial Temperature 25C Initial Pressure 0.099MPa Negative sensitivity indicates a drop in fitness relative to the tuned ring-dynamic model. +15%, TDC +15%, Full Cycle -15%, TDC -15%, Full Cycle -75 -60 -45 -30 -15 0 15 Percent Sensitivity Figure A2 – Sensitivity of crevice modeling parameters based on fitness definition. TDC and full cycle fitnesses shown. provides a good combination to match the experimental data. Additionally, some parameters are more sensitive at either TDC or at the end of the experiment. An example is the value of the circumferential area of the elastomeric ring, A35c; this is more critical at the end of the experiment where this determines how much blowby there is past the ringpack. The most critical parameters in the model are the primary flow path (A13c) and the primary volumes that capture the crevice charge (vol3 & vol4). There is no sensitivity for ct2 since this clearance height is set large enough so that volumes 3 and 4 are in pressure equilibrium (as was seen in Figs. 10 and 15). It can be seen that most changes result in a drop in the fitness parameter, indicating that the set of crevice model parameters chosen for the ringpack model Author:Gilligan-SID:12426-GUID:27843746-134.48.93.192