Quenching steels with gas jet arrays

Similar documents
Simulation of Flameless Combustion of Natural Gas in a Laboratory Scale Furnace

CFD ANALYSIS OF CONVECTIVE FLOW IN A SOLAR DOMESTIC HOT WATER STORAGE TANK

CFD Analysis for Production of Carbon Nanotubes

INVESTIGATION OF FLOW CHARACTERISTICS OF HELLER-TYPE COOLING TOWERS WITH DIFFERENT COOLING DELTA ANGLES

COMBINED EXPERIMENTAL AND SIMULATION (CFD) ANALYSIS ON PERFORMANCE OF A HORIZONTAL TUBE REACTOR USED TO PRODUCE CARBON NANOTUBES

Dr. J. Wolters. FZJ-ZAT-379 January Forschungszentrum Jülich GmbH, FZJ

Heat transfer enhancement in fire tube boiler using hellically ribbed tubes

FEATURE Vacuum/Surface Treating

Gas Quenching With Air Products Rapid Gas Quenching Gas Mixture

Numerical Simulation on Effects of Electromagnetic Force on the Centrifugal Casting Process of High Speed Steel Roll

Performance of a Vertical Axis Wind Turbine under Accelerating and Decelerating Flows

CFD and Wind Tunnel Study of the Performance of a Multi- Directional Wind Tower with Heat Transfer Devices

Heat transfer modelling of slot jet impinging on an inclined plate

Investigating Two Configurations of a Heat Exchanger in an Indirect Heating Integrated Collector Storage Solar Water Heating System

Numerical analysis of eccentric orifice plate using ANSYS Fluent software

Modelling of Material Removal in Abrasive Flow Machining Process Using CFD Simulation

Investigating two configurations of a heat exchanger in an Indirect Heating Integrated Collector Storage Solar Water Heating System (IHICSSWHS)

GAS QUENCHING WITH CONTROLLABLE HEAT EXTRACTION

The 3-D Numerical Simulation of a Walking Beam Type Slab Heating Furnace with Regenerative Burners

Performance Analysis for Natural Draught Cooling Tower & Chimney through Numerical Simulation

Development of Micro Combustion Systems: Insights through Computations and Experiments

Computational Analysis Of Ejector With Oscillating Nozzle

CFD tools and lower order modeling for regenerative chambers with gas recirculation system

CFD SIMULATION AND EXPERIMENTAL VALIDATION OF FLUID FLOW IN LIQUID DISTRIBUTORS

MODELLING COMBUSTION AND THERMAL NO X FORMATION IN ELECTRIC ARC FURNACES FOR THE PRODUCTION OF FERRO-SILICON AND SILICON-METAL

Study on the Cooling Capacity of Different Quenchant

Numerical and Experimental Modeling of Producer Gas Carburettor

Performance estimation on micro gas turbine plant recuperator

Simulation of Melting Process of a Phase Change Material (PCM) using ANSYS (Fluent)

Design and distribution of air nozzles in the biomass boiler assembly

By Kimbal A. Hall, PE. Submitted to: WESTFALL MANUFACTURING COMPANY. May ALDEN RESEARCH LABORATORY, INC. 30 Shrewsbury Street Holden, MA 01520

Effect of Rotation Speed of a Rotary Thermal Wheel on Ventilation Supply Rates of Wind Tower System

Experimental study and numerical simulation of flow pattern and heat transfer during steam drying wood

Simulation of Atmospheric Air Micro Plasma Jet for Biomedical Applications

CFD Guides Design of Biomass Boiler Retrofit to Increase Capacity by 25% and Decrease Ash Carryover by 60%

Brazed aluminium heat exchangers (BAHXs), also referred to

MODERN PRACTICES FOR MEASUREMENT OF GAS PATH PRESSURES AND TEMPERATURES FOR PERFORMANCE ASSESSMENT OF AN AXIAL TURBINE

CFD ANALYSIS OF SOLAR HEATER WATER PIPE WITH DIFFERENT INCLINATION

Numerical Analysis of Two-phase Flow and Heat Transfer in External Cyclic Purification Hot-dip Galvanizing Bath

Evaluating Performance of Steam Turbine using CFD

UTC R170. Hydrogen Flammability Limits and Implications on Fire Safety of Transportation Vehicles. Umit O. Koylu

COOLING OF ROLLS USED IN HOT ROLLING OF LONG PRODUCTS

Application of CFD for Analyzing Quenching and Quench System Design. Andrew L. Banka

ADVANCES in NATURAL and APPLIED SCIENCES

DEVELOPMENT OF THE TUBE QUENCHING PROTOTYPE UNIT

Comparison between Gas and Oil Injection into the Blast Furnace: A Detailed Theoretical Study

CRHT VII. Design and CFD analysis of Pico- hydro Turgo turbine. Paper no. CRHT17-11

Computational Modeling of Counter Flow Heat Exchanger for LMTD Analysis

Project Background (long term goals)

ADVANCES in NATURAL and APPLIED SCIENCES

SIMULATION OF SOLIDIFICATION PROCESS USED IN TWIN ROLL CASTING

Numerical method for modelling spray quenching of cylindrical forgings

EFEECT OF HYDROGEN ADDITION ON METHANE COMBUSTION IN A CAN TYPE COMBUSTOR

Optimising design of secondary combustion chambers using CFD

heat exchanger modelling for generator ventilation systems

CCC Annual Report Quantitative Measurement of Molten Steel Surface Velocity with Nail Boards and SVC Sensor in CC Molds

BOUNDARY SETTING IN SIMULATING GREENHOUSE VENTILATION BY FLUENT SOFTWARE

SCORG TM Setup for CFD Simulation of Twin Screw Machines with STAR-CCM+

E APPENDIX. The following problems are intended for solution using finite element. Problems for Computer Solution E.1 CHAPTER 3

GMAW SHIELDING GAS FLOW OPTIMISATION BY REFINEMENT OF NOZZLE GEOMETRY

Three-Dimensional Numerical Simulation of a Model Wind Turbine

Analysis of Flue Gas Flow Behavior in Economiser Duct Using Cfd

40-Ton Articulated Truck Cooling System Modelling Using STAR-CCM+

Numerical Modeling of Buoyancy-driven Natural Ventilation in a Simple Three Storey Atrium Building

Large Eddy Simulation of temperature distribution in an aero engine annular combustor with a swirler and Nanoparticle injection

5 th OpenFOAM Workshop, Gothenburg, Sweden, June 21-24th, 2010 Using OpenFOAM for Tunnel Ventilation Design

Flow Patterns and Thermal Behaviour in a Large Refrigerated Store

COMPUTATIONAL FLUID DYNAMIC COMBUSTION MODELLING OF A BAGASSE BOILER

Computational Analyses of Combustive Vortex Flows in Liquid Rocket Engines

INFLUENCE OF HEAT SOURCE LOCATION ON AIR TEMPERATURES IN SEALED MV SWITCHGEAR

Analysis of Uniformity and Energy Consumption in. Supermarket installed DurkeeSox Air Dispersion System. with PE Air Dispersion Model

Modelling of plasma heating of the substrate

MODELING OF CHAR COMBUSTION IN CO 2 /O 2 AND N 2 /O 2 ATMOSPHERES

Modeling of Temperature Distribution in Metalcutting using Finite Element Method

Thermal Management of Densely-packed EV Battery Set

Numerical modelling of shielding gas flow and heat transfer in laser welding process

CFD Analysis of Earth-Air Heat Exchanger to Evaluate the Effect of Parameters on Its Performance

Liquid-Solid Phase Change Modeling Using Fluent. Anirudh and Joseph Lam 2002 Fluent Users Group Meeting

CFD Simulation Studies on Integrated Approach of Solar Chimney and Earth Air Tunnel Heat Exchanger for Building Space Conditioning

A computational model for the cooling phase of injection moulding. Abstract

Example 8 - Hopkinson Bar

Flow visualization at suction of a twin screw compressor

Performance Evaluation of Natural Ventilation Devices to Improve Thermal Comfort of a Computer Lab of University Building Using CFD as a Tool

FINITE VOLUME ANALYSIS OF TWO-STAGE FORGING PROCESS FOR ALUMINIUM 7075 ALLOY

Code No: RR Set No. 1

COMSOL Multiphysics Simulation of Flow in a Radial Flow Fixed Bed Reactor (RFBR)

On the performance of the Stravent ventilation system in an office space Numerical and experimental investigations

Investigation on Core Downward Flow by a Passive Residual Heat Removal System of Research Reactor

Available online at ScienceDirect. Procedia Engineering 105 (2015 )

Phoenix Temperature Measurement. Surface Engineering and Heat Treatment Industry Conference. 16 October 2015 Stratford-upon-Avon

Impellers of low specific speed centrifugal pump based on the draughting technology

CFD/FEM Based Analysis Framework for Wind Effects on Tall Buildings in Urban Areas

CERTIFICATES OF COMPETENCY IN THE MERCHANT NAVY MARINE ENGINEER OFFICER

Heat Transfer in Laser Tumor Excision

Heat Storage Performance of a Honeycomb Ceramic Monolith

IMPURITIES REMOVAL PROCESS FROM THE VACUUM INDUCTION FURNACE CHARGE: A NUMERICAL STUDY

Fundamentals of Metal Forming

Literature Review on CFD Study of Producer Gas Carburetor

COLD FLOW SIMULATION OF QUENCHING MEDIA IN AGITATED QUENCH TANK WITH DIFFERENT CONFIGURATIONS USING CFD SOFTWARE

Computer Models For Fire and Smoke

Transcription:

Quenching steels with gas jet arrays PAUL F STRATTON ANDREW P RICHARDSON BOC Rother Valley Way, Holbrook, Sheffield UNITED KINGDOM Paul.stratton@boc.com http://www.catweb.boc.com Abstract: - Single components heated in vacuum or conventional atmosphere furnaces can be quenching by gas jets at higher quenching rates than those possible with conventional multi-component gas quenching. Such treatments may be seen as meeting the need for a clean, non-toxic quenching medium that leaves no residues to be removed after processing. The processing of single components allows the operator complete control of the quenching intensity both locally and generally. Moreover the quenching rate may be changed during the quenching cycle. In order to achieve the best system performance it is necessary to optimise the performance of the jet array. Two-dimensional CFD was used to model and optimise the distance from jet to workpiece, the distance between jets and the gas velocity, and the mean heat transfer coefficient was calculated. A test rig to quench a gear blank was constructed using the optimised array. This rig was then modelled in three dimensions and the results compared to those generated experimentally. Some modifications to the model were necessary to align it with the experimental results. Keywords: - Quenching, Nitrogen, Jet arrays, Modelling, Validation, Steel 1 Introduction Gas quenching in vacuum furnaces has been used for several years, usually under nitrogen, argon or helium at pressures up to 60 bar, and its characteristics for bulk quenching of components are well known [1]. It has recently been suggested that gas quenching could be applied to single components or small groups that were heated in either vacuum or conventional atmosphere furnaces [2]. To eliminate the need to cool the furnace structure, these techniques often require the transfer of the component to be quenched to a specially designed cold chamber [3]. Gas jet quenching applied to a single component heat treated in either vacuum or conventional atmosphere furnaces can achieve higher quenching rates than would be possible with conventional multi-component gas quenching [4]. Such treatments may be seen as meeting the need for a clean, non-toxic quenching medium that leaves no residues to be removed after processing. The processing of single components using a jet array allows the operator almost complete control of the quenching intensity both locally and generally. Moreover the quenching rate can be changed during the quenching cycle. It may therefore be possible to marquench one area of a component and fast oil quench another in a single operation. If the benefits of gas quenching are to expand beyond the niche process areas where its high uniformity and high repeatability justify a premium price, then a process that uses the minimum quantity of the lowest possible cost gas is required. It is an unfortunate fact that the high conductivity gases such as helium and hydrogen are expensive. The low cost alternative, nitrogen, has poor thermal performance. However, it has been shown that the delivery pressure available from liquid nitrogen storage systems is able to generate the jet velocities required to produce at least oil-like quenching characteristics [4]. 2 Optimising the jet array In order to meet the criteria for uniform quenching of a single component, the quenchant has to cool the surface of the component uniformly. For a gas to achieve this aim, it must have the same speed and direction at all points although in practice it has to enter through discrete nozzles. Modelling shows that, under most conditions, the cooling is at a maximum directly in line with the nozzle and falls away to the mid-point between the nozzles, as shown in Fig. 1.

Fig. 1. The surface heat transfer coefficient for v=100 m/s, a= 50.8 mm (2 inches) and b=88.9 mm (3.5 inches). The quench process was modelled using the Fluent v5.0 computational fluid dynamics software package. For this initial screening only two dimensions were modelled to speed up the process. The model consisted of an array of gas nozzles 12.7 mm (0.5 inches) in diameter perpendicular to the hot surface. The distance between the nozzles and the surface (a) and the distance between nozzles themselves (b) was varied for a range of imposed gas velocities (v) at the exit from the nozzle. A typical velocity profile derived from the model is shown in Fig. 2. A closer view of the velocities at the surface (Fig. 3) shows that the flow over the major part of the surface is far from the optimal perpendicular and is in fact parallel to it, reducing the maximum heat extraction rate. Fig. 3. A close up view of the velocity profile for v=100 m/s, a= 50.8 mm (2 inches) and b=88.9 mm (3.5 inches). The heat transfer coefficient for the hot surface was calculated as a function of the distance from the centre line of the nozzle. The surface heat transfer profile for each set of conditions was integrated to give the average heat transfer coefficient. These values were plotted as a function of distance between nozzles and the surface (a) in Fig. 4. Fig. 2. A typical velocity profile for v=100 m/s, a= 50.8 mm (2 inches) and b=88.9 mm (3.5 inches). Fig. 4. The variation of mean surface heat transfer coefficient with the distance between the surface and the nozzles (a) for v=100 m/s and b=12.7 mm (0.5 inches)

distance between nozzles of 190.5 mm (7.5 inches) and a gas velocity of 100 m/s. Fig. 5. The variation of mean surface heat transfer coefficient with the distance between the nozzles (b) for v=100 m/s and a=3.2 mm (0.125 inches) To keep down costs it is obviously necessary to minimise gas flow. As the gas flow for a given nozzle is fixed by the cooling rate required, the only variable available is the distance between nozzles. Somewhat surprisingly, the distance between the nozzles has little effect on the heat transfer coefficient as can be seen from Fig. 5. This effect is due to the area of high turbulence created at the edge of the nozzle at high gas velocities. This effect can clearly be seen in the velocity vector diagram (Fig. 6). Fig. 6(a). The velocity vector diagram for a nozzle to surface distance of 3.2 mm (0.125 inches), distance between nozzles of 190.5 mm (7.5 inches) and a gas velocity of 100 m/s. Fig. 6(b). A detail close to the nozzle edge from the velocity vector diagram for a nozzle to surface distance of 3.2 mm (0.125 inches), The heat transfer coefficient is also relatively insensitive to scaling factors; i.e. if all the size factors are reduced by a factor of four, which is likely to include the maximum practical range of jet sizes, there is only a 30% increase in the heat transfer coefficient [5]. This lack of sensitivity to the size of the nozzles and the distance between them simplifies the design of quenching enclosures, especially for complex shapes. However the close approach to the surface required by the technique does result in the need for careful consideration of the nozzle sites. As a result of the high gas pressures, it should be possible to eliminate the need to support the product during quenching. The effect of the product s weight will be small compared with the applied force of the gas and the product would float within the nozzle field. If the ratio of the diameter of the nozzle and the distance between the nozzle and the surface is chosen as four (the point at which the area for gas escape equals the area of the nozzle), then the system is selfcompensating. This is because any reduction in distance will increase the pressure at the nozzle for a given flow and increase the separation again. Small inconsistencies would be introduced into the flow field in a practical device and would lead to oscillation or rotation of the component, which would produce more even quenching. The high velocities used will probably lead to high noise levels in the vicinity of the quench, but this effect could be minimised by sound insulation around the cold wall quenching chamber. Mean heat transfer coefficient (W/m 2 0 K) 1600 1400 1200 1000 800 600 400 200 0 0 50 100 150 200 250 300 Gas velocity (m/s) Fig. 7. The effect of velocity on mean heat transfer coefficient for a=3.2 mm (0.125 inches) and b=38 mm (1.5 inches) The cooling rate is almost directly proportional to the gas velocity at gas velocities below 100 m/s (Fig. 7), and the velocity is related to the supply pressure. It is

obviously simple, therefore, to control the cooling rate. Although very high velocities towards sonic will result in higher cooling rates, the rate of increase becomes non-linear, so their use is likely to be restricted to applications where the highest possible cooling rates are required. 3 Validating the model In order to validate the model it was necessary to construct a test rig. When the rig had been constructed it was modelled with Fluent 6.0.12 using the same conditions that had been used in earlier work (Stratton et al, 2000), except that some of the physical characteristics of the steel were replaced with more recent data. Fig. 8. The model Initially the standard Fluent segregated solver was used together with first order discretisation for momentum, energy and turbulence parameters; nitrogen was simulated as an incompressible gas. The k-e turbulence model was also used with standard wall functions. An initial steady solution was achieved for a cold flow and then for a fixed hot sample temperature. During this period the mesh at the tips of the nozzles and on the surface of the sample were repeatedly adapted to ensure that the boundary layer assumptions of the model were within the valid range (Y + in the range 30-60). The solver was then switched to the unsteady mode (1 st order implicit) with a time step of 1 second. Although the maximum number of iterations per second was set to 100 to ensure a good degree of convergence at each time step, the model would typically converge with considerably fewer iterations, especially towards the end of the simulation. Convergence criteria were set for normalized unscaled residuals of 10-3 for continuity, velocity and turbulence and 10-6 for energy and radiation, with model mass and energy imbalances monitored periodically across boundaries. Calculations performed on a dual 1.7 GHz P4 processor Dell Precision workstation with 1 GB RAM took approximately 90 seconds per iteration. Because of time constraints and the close correlation with experimental results this work did not examine at the effects of various modeling options, e.g. higher order discretisation schemes, turbulence models and compressibility. The model domain was set at a 200 mm radius, almost twice the radius of the specimen, and extended vertically to the gas distribution manifolds above and below the sample (200 mm total). The specimen and tube bank arrays to form the jets were thus centrally located within this cylindrical domain (Fig. 8). The model was meshed in two parts. The internals of the gas feed tubes and specimen itself were meshed using a regular hexahedral scheme (50,000 cells), while the gas space was meshed using a pyramidal scheme (800,000 cells). Attention was placed particularly on resolving the mesh near the tube tips and specimen surface. Because different mesh densities were needed in the specimen and in the gas flow space, nonconformal interfaces were set up on the specimen boundaries. This allowed higher quality meshes with fewer cells to be generated for each region. Fig. 9. The temperature distribution after 2 seconds

respectively. The model was also used to predict the cooling curve at the position of the thermocouple. Several experimental runs were carried out. The results (Fig. 13) show clearly that the model predicted a much lower initial cooling rate than occurred in practice. Fig. 14 compares the actual cooling rate with that predicted by the model. The large difference in initial cooling rate was probably caused by radiative heat losses that were not taken into consideration in the model. Fig. 10. The temperature distribution after 6 seconds Fig. 13. Actual and predicted cooling curves Fig. 11. The temperature distribution after 30 seconds Fig. 14. Comparison of the actual cooling rate with that predicted by the model Fig. 12. The temperature distribution after 86 seconds Typical model outputs for 2, 6, 30 and 86 seconds quenching are shown in Figs. 9 to 12 The modelling was therefore repeated taking radiative heat losses into account. The effect of this modification is shown in Fig. 15. The discreteordinates model was chosen to simulate the radiative losses with a sample surface emissivity of 0.8. The simulation was started from the actual pre-heat furnace temperature to allow for a more representative thermal distribution at the start of the gas quench. This choice of assumed emissivity was validated, as the modelled cooling by radiation was at

a similar rate to that measured in the experiment. There is now a significant difference between the cooling rates at the surface and in the core of the sample, due to the high initial heat loss from the surface by radiation. The modelled cooling rate in the core is increased over the same sort of temperature range as the actual but not to the same magnitude. Fig. 15. Comparison between actual and modelled cooling rates when radiation is taken into account It was suggested that the equilibrium specific heat capacity (Cp) values used for the model were inappropriate for continuous cooling conditions [6]. Cp values appropriate to the phases present at the time derived from the continuous cooling curve were substituted with the results shown in Fig. 16. This gave an almost perfect match with the experimental data. using CFD. An array with the jets about four to eight times their own diameter apart and the distance of a quarter of the diameter from the surface with a jet velocity of 100 m/s was found to be optimum. The results of experimental work suggested that radiative heat loses must be included in the model. It also showed the importance of using dynamic rather than equilibrium data for modelling steels under a rapidly changing temperature regime. References: 1. P.F. Stratton, N. Saxena and R. Jain; Requirements for Gas Quenching Systems, Heat Treatment of Metals, 1997.3, 1997, pp. 60-63. 2. B Edenhofer; An overview of advances in atmosphere and vacuum heat treatment, Heat Treatment of Metals, 1999.1, 1999, pp. 1-5. 3. F.T. Hoffmann, T. Lübben, and P. Mayr, Innovations in quenching systems and equipment: current status and future developments, Heat Treatment of Metals, 1999.3, 1999, pp. 63-67. 4. P F Stratton, D Ho and N Saxena Modelling of high speed gas quenching, Journal of Shanghai Jiaotong University, Vol. E-5, No. 1, 2000, pp146-155. 5. P F Stratton and D Ho Individual component gas quenching, Heat Treatment of Metals, Vol.28, No.3, 2000, pp 65-68. 6. Private communication, S. Segerberg, IVF, Sweden (2002). Fig. 16. Comparison of the original and revised model quenching curves with the experimental data 4 Conclusions Jet arrays for quenching steel components can successfully be optimised by modelling them