A MODEL FOR BINDER GAS GENERATION AND TRANSPORT IN SAND CORES AND MOLDS

Similar documents
FLOW-3D Core Gas Model: Binder Gas Generation and Transport in Sand Cores and Molds. A. J. Starobin and C. W. Hirt Flow Science, Inc.

GAS PRESSURE IN ALUMINUM BLOCK WATER JACKET CORES. Abstract

Characterizing Flow Losses Occurring in Air Vents and Ejector Pins in High Pressure Die Castings

SIMULATION OF ENTRAPMENTS IN LCM PROCESSES

CONSIDERATION OF DENSITY VARIATIONS IN THE DESIGN OF A VENTILATION SYSTEM FOR ROAD TUNNELS

Numerical Simulations of a Train of Air Bubbles Rising Through Stagnant Water

Experimental Characterization and Modeling of Helium Dispersion in a ¼-Scale Two-Car Residential Garage

Hydrostatics Physics Lab XI

A Computational Assessment of Gas Jets in a Bubbly Co-Flow 1

CFD SIMULATIONS OF GAS DISPERSION IN VENTILATED ROOMS

MODELING OF THERMAL BEHAVIOR INSIDE A BUBBLE

WP2 Fire test for toxicity of fire effluents

PHYS 101 Previous Exam Problems

LOW PRESSURE EFFUSION OF GASES adapted by Luke Hanley and Mike Trenary

You should be able to: Describe Equipment Barometer Manometer. 5.1 Pressure Read and outline 5.1 Define Barometer

Name Chemistry Pre-AP

End of Chapter Exercises

LOW PRESSURE EFFUSION OF GASES revised by Igor Bolotin 03/05/12

Permeability. Darcy's Law

Old-Exam.Questions-Ch-14 T072 T071

HYDROGEN RISK ANALYSIS FOR A GENERIC NUCLEAR CONTAINMENT VENTILATION SYSTEM

MODELING&SIMULATION EXTRAPOLATED INTERNAL ARC TEST RESULTS: A COUPLED FLUID-STRUCTURAL TRANSIENT METHODOLOGY

Injector Dynamics Assumptions and their Impact on Predicting Cavitation and Performance

ANALYSIS OF HEAT TRANSFER THROUGH EXTERNAL FINS USING CFD TOOL

Offshore platforms survivability to underwater explosions: part I

NUMERICAL INVESTIGATION OF THE FLOW BEHAVIOUR IN A MODERN TRAFFIC TUNNEL IN CASE OF FIRE INCIDENT

Fall 2004 Homework Problem Set 9 Due Wednesday, November 24, at start of class

Chapter 13 Temperature, Kinetic Theory, and the Gas Laws 497

Development of Fluid-Structure Interaction Program for the Mercury Target

Computer Simulation Helps Improve Vertical Column Induced Gas Flotation (IGF) System

Accurate Measurement of Steam Flow Properties

Traffic circles. February 9, 2009

Study on the Influencing Factors of Gas Mixing Length in Nitrogen Displacement of Gas Pipeline Kun Huang 1,a Yan Xian 2,b Kunrong Shen 3,c

Unit 2 Kinetic Theory, Heat, and Thermodynamics: 2.A.1 Problems Temperature and Heat Sections of your book.

End of Chapter Exercises

CFD Simulation and Experimental Validation of a Diaphragm Pressure Wave Generator

AIRFLOW GENERATION IN A TUNNEL USING A SACCARDO VENTILATION SYSTEM AGAINST THE BUOYANCY EFFECT PRODUCED BY A FIRE

Chapter 2 Hydrostatics and Control

. In an elevator accelerating upward (A) both the elevator accelerating upward (B) the first is equations are valid

PREDICTION OF TOTAL PRESSURE CHARACTERISTICS IN THE SETTLING CHAMBER OF A SUPERSONIC BLOWDOWN WIND TUNNEL

Flow and Mixing in the Liquid between Bubbles

Gas Pressure. Pressure is the force exerted per unit area by gas molecules as they strike the surfaces around them.

Golf Ball Impact: Material Characterization and Transient Simulation

Huntsman Polyurethanes smart simulation software. Process optimization by simulation

S.A. Klein and G.F. Nellis Cambridge University Press, 2011

To convert to millimeters of mercury, we derive a unit factor related to the equivalent relationship 29.9 in. Hg = 760 mm Hg.

Micro Channel Recuperator for a Reverse Brayton Cycle Cryocooler

Analysis of a KClO3 Mixture and Determination of R

Fluid-Structure Interaction Analysis of a Flow Control Device

Blast Damage Consideratons for Horizontal Pressure Vessel and Potential for Domino Effects

ZIN Technologies PHi Engineering Support. PHi-RPT CFD Analysis of Large Bubble Mixing. June 26, 2006

U S F O S B u o y a n c y And Hydrodynamic M a s s

ISOLATION OF NON-HYDROSTATIC REGIONS WITHIN A BASIN

A Numerical Simulation of Fluid-Structure Interaction for Refrigerator Compressors Suction and Exhaust System Performance Analysis

A Model of Gas Bubble Growth by Comsol Multiphysics

Bio-Effluents Tracing in Ventilated Aircraft Cabins

EXPERIMENTAL STUDY ON THE HYDRODYNAMIC BEHAVIORS OF TWO CONCENTRIC CYLINDERS

Flow transients in multiphase pipelines

The University of Hong Kong Department of Physics Experimental Physics Laboratory

SIMULATIONS OF HYDROGEN RELEASES FROM A STORAGE TANKS: DISPERSION AND CONSEQUENCES OF IGNITION

1 PIPESYS Application

Questions. theonlinephysicstutor.com. facebook.com/theonlinephysicstutor. Name: Edexcel Drag Viscosity. Questions. Date: Time: Total marks available:

Thermodynamic Study of Compartment Venting

MODELING STUDY OF THE BEHAVIOR OF LIQUID FIRE SUPPRESSION AGENTS IN A SIMULATED ENGINE NACELLE

CHEMISTRY - CLUTCH CH.5 - GASES.

DUE TO EXTERNAL FORCES

Domain Decomposition Method for 3-Dimensional Simulation of the Piston Cylinder Section of a Hermetic Reciprocating Compressor

SOLAR 93 THE 1993 AMERICAN SOLAR ENERGY SOCIETY ANNUAL CONFERENCE. Washington, DC April 22028,1993. Editors: S. M. Burley M. E.

States of Matter. Q 7. Calculate the average of kinetic energy, in joules of the molecules in 8.0 g of methane at 27 o C. (IIT JEE Marks)

CHEM 3351 Physical Chemistry I, Fall 2017

Workshop 1: Bubbly Flow in a Rectangular Bubble Column. Multiphase Flow Modeling In ANSYS CFX Release ANSYS, Inc. WS1-1 Release 14.

An underwater explosion is an explosion where the point of detonation is below the surface of the water.

Generating Calibration Gas Standards

An Innovative Solution for Water Bottling Using PET

Model of Gas Flow through Porous Refractory Applied to An Upper Tundish Nozzle. Rui Liu and Brian G. Thomas

Analysis of Pressure Rise During Internal Arc Faults in Switchgear

MATCHING EXPERIMENTAL SATURATION PROFILES BY NUMERICAL SIMULATION OF COMBINED AND COUNTER-CURRENT SPONTANEOUS IMBIBITION

LAB 13: FLUIDS OBJECTIVES

EXPERIMENTAL STUDY ON BEESWAX USING WATERJET DRILLING

Increase in Evaporation Caused by Running Spa Jets swhim.com

Experimental study on path instability of rising bubbles

Monitoring Gas Pressure in Vacuum Insulation Panels

Honors Chemistry - Problem Set Chapter 13 Classify each of these statements as always true, AT; sometimes true, ST; or never true, NT.

Rotary air valves used for material feed and explosion protection are required to meet the criteria of NFPA 69 (2014)

EFFECTIVE DESIGN OF CONVERTER HOODS. 111 Ferguson Ct. Suite 103 Irving, Texas U.S.A. 400 Carlingview Dr. Toronto, ON M9W 5X9 Canada.

Novel empirical correlations for estimation of bubble point pressure, saturated viscosity and gas solubility of crude oils

Impact of imperfect sealing on the flow measurement of natural gas by orifice plates

MODELLING OF FUME EXTRACTORS C. R.

Detector Carrier Gas Comments Detector anode purge or reference gas. Electron Capture Nitrogen Maximum sensitivity Nitrogen Argon/Methane

THERMODYNAMICS OF A GAS PHASE REACTION: DISSOCIATION OF N 2 O 4

White Paper. Chemical Sensor vs NDIR - Overview: NDIR Technology:

Boyle s Law Practice

EXPERIMENT 8 Ideal Gas Law: Molecular Weight of a Vapor

Figure Vapor-liquid equilibrium for a binary mixture. The dashed lines show the equilibrium compositions.

Visual Observation of Nucleate Boiling and Sliding Phenomena of Boiling Bubbles on a Horizontal Tube Heater

A New Way to Handle Changing Fluid Viscosity and the Full-to-empty Effect

Runs Solution Temp. Pressure zero reset

Air Bubble Defects in Dispensing Nanoimprint Lithography

Accelerated Chemistry Study Guide Chapter 13: Gases

AP TOPIC 6: Gases. Revised August General properties and kinetic theory

Transcription:

Modeling of Casting, Welding, and Solidification Processes XII TMS (The Minerals, Metals & Materials Society), 2009 A MODEL FOR BINDER GAS GENERATION AND TRANSPORT IN SAND CORES AND MOLDS Andrei Starobin 1, C.W. Hirt 1, D. Goettsch 2 1 Flow Science, Inc.; 683A Harkle Road; Santa Fe, New Mexico, 87505, USA 2 GM Powertrain, 777 Joslyn Ave; Pontiac, Michigan, 48340, USA Keywords: Computational fluid dynamics; Fixed-mesh method; Thermal degradation of binders; Core Gas Flow; Casting Abstract Gas defects are often found in sand-cast parts that arise from gas generated by thermal decomposition of the sand binder. Typically, these defects occur with sand cores that have not been adequately vented. A new computational model has been developed that predicts the generation of decomposed binder gas and its transport within cores. The binder gas is treated as compressible to account for poor venting scenarios and for transport across core regions of steep temperature gradients. The model considers true molding geometry and actual core venting locations and can be used to predict the amount of gas that would enter liquid metal from any location on the surface of the core. Overview The making of resin-bonded sand castings has made great strides in quality over its long history. Even so, there remain some process-related defects that are not fully understood and can cause quality issues. For instance, chemical binders in the sand can produce gas when heated by the molten metal and if this gas is not vented adequately, the gas may flow into the metal resulting in a gas porosity defect. This is most likely with cores that form thin interior features of castings that heat up quickly and have long venting paths. In addition to the core geometry, the composition of the binder may play an important role as well. Two major types of binders are used in core making practice: resin-based organic binders and inorganic binders such as sodium silicate [9]. The organic binders are either thermosetting, or cured at room temperature with an aid of a catalyst. These are favored in many applications due to their complete degradation even at aluminum casting temperatures and for the ease of subsequent sand shake out. The presented core gas model is developed with these binders in mind, but can be extended to inorganic binders if appropriate data on their decomposition is available. The thermal degradation of a resin-based binder can be experimentally characterized in two ways: by measuring the mass loss of the binder over time at a fixed heating rate [3,4, 13], or by measuring the amounts and types of volatile and condensable decomposition products over time and over a range of temperatures [5,7, 10, 11]. A reliable model for core gas flow which would predict core gas pressures requires both types of data. The experimental techniques that quantify the amount of binder gas can be broadly classified as displacement in type [7, 10], or as pressure based methods [11]. In the former, collected gas displaces oil in a tank with the back of the tank vented to the ambient pressure. In the latter, the gas is collected in a fixed volume apparatus and the gas pressure is measured. In both cases the

collection lines are kept at an elevated temperature to prevent condensation of the volatile components which contribute to core pressure. Also, in both cases the standard volumetric rate of the product gas is reported over time. The data obtained by the pressure based method was used to calibrate the proposed binder decomposition model. The model assumes that the binder converts completely into gas. The gas is thought to be ideal and of fixed composition with the gas constant, R g. The gas constant is deduced from the total collected standard volume, V std, and the initial mass of the binder, m b : R cg p T = std std (1) std V m b The major factors contributing to core gas pressure are described by Campbell [1]. Briefly, the hot product gas forms over a range of temperatures estimated in this work to be roughly between 500 and 700 K for resin-based binders. The gas then flows toward the vents through porous sand with an associated pressure loss. The sand porosity can be as high as 40% [9]. Venting either occurs naturally at the back of the mold, or is introduced by drilling through the mold halves to the core print locations. The vent locations are near room temperature during the casting pour. Three regions of flow can be identified: the high temperature decomposition region, the flow 5 through a steep thermal gradient which can be as high as 10 K/m on the core surfaces in iron castings, and finally near-isothermal flow away from core surface to the vents. In what follows, we formulate the binder decomposition and mass transport model through porous cores or molds. We then show that a simple two-parameter decomposition model adequately captures volume evolution data from a typical polyurethane cold-box (PUCB) calibration core immersed in iron. The model is the then used to compute thermal expansion driven air flow out of a spherical core which was studied in recent experiments [7]. Finally, we present some validation results for vented and unvented aluminum engine block PUCB jacket cores studied at a General Motors Research Foundry. The simulations demonstrate how our coregas model can be used concurrently with an existing simulation program [12] which is routinely used to study mould filling and casting freezing. The Core Gas Model The microscopic velocity of the core gas, u r cg, is governed by the equation for flow in porous media: r K ucg = pcg, (2) µ 10 where K is the intrinsic sand permeability found to be approximately 10 m 2 in room temperature air flow experiments [11], µ is core gas viscosity and p cg is the core gas pressure. Both the macroscopic inertial terms and the Forchheimer term are found to be small for the typical conditions of the core gas flow and are not included in Equation (2). Furthermore, the possible compositional and temperature dependence of the gas viscosity is not accounted for in the model. The density of the core gas is governed by the mass transport equation:

ρ cg r d ρ b + ( ρ cgucg ) =, (3) t dt where ρ cg and ρ b are microscopic core gas and macroscopic core binder densities. The core gas is compressible. For example, even in the absence of gas sources there can be a thermal expansion and flow of the initial air in the core as the core heats up. Furthermore, while the pressures in the cores are expected to be a few psi above ambient and are effectively limited by the surrounding metal pressure, the temperature variations from core surface to the vents are factors of two suggesting significant contraction of core gas during flow. The density transport equation, Equation (3), must be numerically solved with an implicit coupling to the velocity equation to insure computational stability at reasonable time-step sizes. This is achieved by using a variant of the Implicit Continuous-fluid Eulerian (ICE) method [6]. The core gas density is further constrained by the ideal gas law which should describe well the binder product gas, especially at high temperatures: p = R ρ T (4) cg cg cg The gas constant Rcg is computed using Equation (1), while the temperature of the gas is assumed to be the local core temperature. This is a good approximation because of the high heat content of the solid core material compared to that of the gas. Since it is assumed that the core and gas temperatures are equal, and the core temperature T is already computed by a heat transfer solver, the ideal gas law can be used to compute gas pressure when its density is given or vice versa. In the present model the core gas density is evaluated from a transport Equation (3). The conversion of the solid binder to gas is described by an Arrhenius relationship in accord with a number of previous thermal decomposition studies on polymers [3, 4, 13]: d ρ b Eb = ρ bcb exp, (5) dt RT Where ρ b is the solid binder density, C b is an empirical reaction rate constant, Eb is a component binding energy, R is the universal gas constant and T is the core temperature. Study [4] is particularly relevant since the two reaction constants were determined for pyrolisis of 10 1 polyurethane resins at different heating rates. The two constants are 134 kj/mol and 4.2510 s and the reaction order is close to one hence the simple form of Equation (5). The highest heating rate probed in this pyrolisis study was only 0.3 deg/sec, while at the core surface in iron immersions the heating rate is as high as 80 deg/sec. However, at this point this is the best data available and we use it for the simulations described below. Using the exponential Arrhenius rate for binder decomposition can introduce some undesirable computational effects. For example, for the decomposition law of Equation (5) with the listed reaction constants, most of the binder loss happens between 645 and 700 K. Given the steep 5 temperature gradient of about10 K/m at the core surface, even at mesh resolutions of 2 mm the change in temperature between the surface core cell and its neighbor in the core interior is larger than the decomposition zone width. Thus at this mesh resolution the source may exhibit an unphysical oscillatory behavior associated with the discrete size of the computational elements. To alleviate this problem simple subdivision scheme is used in each computational cell. In the

sub-cells solid binder density is stored, while the sub-cell temperature is obtained through linear interpolation of the coarse temperature field. Two more simplifying assumptions are made at this point. First, no gas is allowed to condense within the core. For most practical cases, condensation, if it occurs, is not likely to significantly affect the early gas pressures that are the principal source for gas related casting defects. Second, the possible endo- or exo-thermic effects of pyrolitic binder decomposition are ignored. Because the core-gas model is used concurrently with a mold filling model [12], it is possible that gas flow in cores may become sufficiently fast to limit the computational time-step size (by the Courant stability condition for an explicit advection approximation) to a value smaller than what is needed for the filling simulation, resulting in longer calculation times. To counter this possibility, a sub-time-stepping scheme has been incorporated into the core-gas model. This is possible because the model is coupled to the metal filling simulation through boundary conditions (see below). If it is found that the time-step size for the core-gas model must be smaller than that for the filling simulation, then the time-step size for the core-gas computations is reduced to a stable value. The smaller value is always some integer fraction of the original value so that the gas computations can be repeated for that integer number of steps to bring the solution forward to the correct time. The core-gas computations are much faster compared to those for the filling because the gas dynamics is much simpler and the core gas region is only a small portion of a full simulation domain. Core Boundary Conditions At boundaries of the core material there may be a flow of gas either in or out of the core. This exchange is treated as a boundary condition for the core-gas model. The passage of gas through the boundary depends on what is located outside the core. For instance, if the core surface is exposed to air, then gas may flow across the boundary in either direction depending on the pressure gradient. If there is liquid metal at the core surface, then gas is allowed to pass out of the core when its pressure is greater than the pressure of the metal at that location, but no metal is allowed to enter the core. Of course, if the metal has already solidified at the surface of the core, then no gas is allowed to flow across the boundary at that location. In the basic operation of the model, the core gas does not influence metal dynamics. Instead this gas flow is accumulated to identify regions where core gas would percolate into the metal. This option requires the least computational time. Another boundary condition occurs at print surfaces, that is, where a core surface is in contact with another solid part of the mold. At these surface locations, gas does not normally flow unless channels have been cut into the mold to allow for venting. The core-gas model has an option for allowing venting at print surfaces. Core Gas Flow in Iron Calibration Experiments The calibration experiments on small cores were carried out at AlchemCast [11]. A standard size cylindrical core 1.125 inches in diameter and 2 inches long was inserted into an insulating holder and then immersed into the crucible with metal. The size was chosen so that the core gas was sealed at all times with the only escape path at the core top which led into a fixed-volume collection apparatus. The pressure was monitored and subsequently converted into the standard volume.

The left panel of Figure 1 shows the AlchemCast data for a 1.5 wt % PUCB bound core. The gas constant evaluated according to Equation (1) is 357 J/K/mol with a corresponding molar mass of 23 g/mol. This is consistent with the measurements on gas compositions from mold-metal interfaces [5, 9] which indicate that hydrogen, carbon monoxide, methane and carbon dioxide dominate the volatiles (listed in the order of observed abundance). Overall, the agreement is good, with the peak time captured accurately and the peak rate somewhat underestimated. There is some evidence [5, 9] that composition of collected gas varies from light to heavy during the casting process, the inclusion of which may further improve the agreement between data and simulation. However variable gas composition is beyond the scope of the current model and is a possible future extension. The right panel of Figure 1 shows the computed decomposition front geometry for the PUCB calibration core at 30 seconds after immersion. The color variable is the solid binder loss rate 3 which is typically ~ 10 kg/ m /s at, or near the surface in iron immersions. The zone adjacent to the core surface behind the decomposition front contains no intact binder, while the zone adjacent to the vented face has not begun to decompose. Figure 1. Left panel: Calibration data for a 1.125 in. dia. x 2 in. PUCB core immersed in Iron (2600 F) and results from the FLOW-3D core gas model. Right panel: The shape of the binder 3 degradation front at 30 seconds after core immersion. The density loss rate has units of kg/ m /s Air Flow From A Spherical Core Recent experiments done at Swecast [7] on small 6 cm diameter spherical cores, show qualitatively similar volumetric rate curves to the one shown in Figure 1. These measurements were performed with the COGAS apparatus (see [7] for more details) which is based on a displacement technique described above. In [7] the authors point out that the air displaced from the core during immersion is also a contributor to the core pressures. Since our model can be used to study flow of any gas through the porous cores we evaluate the contribution of air to the total volumetric rate signal and the core gas pressure. Our results are summarized in Figure 2. The peak rate of air flow is approximately 0.5 cc/sec and is at about one second after immersion. The peak pressures are only about 5 Pa about the ambient pressure. The experimentally observed peak volumetric rate in [7] is about 20 cc/sec. It is therefore unlikely that air contributes significantly to core pressures in this spherical core, or in larger industrial cores.

Figure 2. Left: Flow field and pressures in a spherical 6 cm dia. core used in [7] to experimentally monitor volume of binder degradation products. The core is held in place against the immersion tube by buoyancy. Right: Simulated flow rate and total standard volume of core air fluxed from the same core Flow in Vented and Unvented Engine Block Jacket Cores Direct observations of core gas bubbles have been made of a vented and an unvented aluminum engine block water jacket core. The jacket and adjoining slab core were placed in an open mold flask, which was bottom filled with 1330F 319 aluminum alloy with a 11 mm/sec vertical velocity (Figure 3). The total fill time was approximately 20 seconds. In the first trial, bubbles were observed through the bulk of the liquid metal as the geometric peaks of the core were submerged. The bubbling stopped when the metal surface reached 75 mm above the top of the cores. A series of vent holes were formed in the second test core piece. Each core print pad had a 3mm diameter hole, 90 mm long. This left a 60 mm distance from the vent hole to the top of the core where the bubbles were observed in the previous test. No bubbles were observed in the submersion of this vented core. Figures 3 and 4 summarize the core-gas model results for the two water jacket core submersions. The differences in geometries between the two jacket cores are clearly seen in the panels of Figure 3. The core on the right shows one of three venting holes. Both the gas pressure and the metal pressure in the mold are plotted. It is clear that the vented core is sealed at this point in the fill (at ~ 18 sec). On the other hand the the unvented core also at 18 seconds shows gas blow. The left panel of Figure 3 plots the gas flux at core surface. Loss of gas into the metal is visible in the top portions of the bottom and top halves of the jacket core. The gas blow occurs at sites of lowest local metallostatic head. Additional, desirable gas venting can be seen at prints of the slab core into the mold flask where perfect venting, fixed pressure boundary conditions are set. The same boundary condition is used in the venting holes of the vented jacket core design. A final observation about the core gas flows in the two cases concerns the peak gas velocity. On average in the vented design the computed velocities are a factor of 2.5 lower. In Figure 4 we plot the cumulative computed core gas mass rate and the gas blow mass rates in the two designs. The vented core shows a significantly lower gas blow rates during the submersion of the bottom portion of the core and a complete suppression of gas blow past 15 seconds. The computed result is in good agreement with the experimental observations, especially later in the fill. The gradual increase of the gas blow from the unvented core is also in

agreement with foundry observations. A row of bubbles (one per cylinder) was produced after the bottom of the jacket core submerged at about 8 seconds and another row after complete core submersion at approximately 15 seconds. Figure 3: Left panel: Locations and amounts of gas flux from the slab and unvented jacket core. Right panel: Metal pressure and gas pressures in a vented jacket core Figure 4: Total binder gas and gas blow into the metal computed for vented and unvented cases. Conclusions We have presented a simple, but predictive, model for binder gas generation and transport in sand cores and molds. Future work will focus on extending the model to capture the effect of residual moisture on gas pressures and on validating the existing model for other binder systems such as shell resin. The presented model can be used to capture gas evolution data from simple cores in iron and steel immersions. In more complicated core design situations, the model can be used to evaluate venting strategies, binder content, sands of different permeability and to compare core and molding designs.

Acknowledgements David Goettsch thanks Jason Traub at the GM Research & Development foundry in Warren MI. C.W. Hirt and Andrei Starobin wish to thank Charles Bates for providing calibration data and Harry Littleton and Michael Barkhudarov for useful discussions. References 1. J. Campbell, Castings (Butterworth-Heinemann, 2000), 105-109 2. Donald A. Nield and Adrian Bejan, Convection in Porous Media (Springer, 2006), 5, 39-40 3. Kalayan Annamalai and Ishwar K. Puri, Combustion Science and Engineering (CRC Press, 2007), 390-391 4. R. Font et al, Pyrolysis study of polyurethane, Journal of Analytical and Applied Pyrolisis, 58-59, (2001), 63-77 5. W.D. Scott and C.E. Bates, Decomposition of resin binders and the relationship between gases formed and the casting surface quality, AFS Transactions, 1980 6. Harlow, F.H. and Amsden, A.A., A Numerical Fluid Dynamics Calculation Method for All Flow Speeds, J. Comput. Phys. 8, 197 (1971) 7. J. Orlenius, U. Gotthardsson and A. Dioszegi, ibid Mould and Core Gas Evolution in Grey Iron Castings, International Journal of Cast Metals Research, 2008, in press 8. Clive Rogers, A virtual tool for the manufacture and use of foundry cores and molds, MCWASP XI, 2006 9. WD Scott, Hanjun Li, John Griffin, and Charles Bates, Production and Inspection of Quality Aluminum Castings, in press 10. L. Winardi, R.D. Griffin, H.E. Littleton, J.A. Griffin, Variables Affecting Gas Evolution Rates and Volumes from Cores in Contact with Molten Metal, AFS Transactions 116, 505-521, 2007 11. AlchemCast LLC, Pelham, Alabama, USA, www.alchemcast.com 12. FLOW-3D Theory Manual, (Flow Science Inc., Santa Fe, New Mexico, USA, www.flow3d.com, 2008). 13. Lengelle, G., Thermal Degradation Kinetics and Surface Pyrolysis of Vinyl Polymers, AIAA J., 8(11), 1989-1196, 1970