Vous êtes sur la page 1sur 14

Analytical/Finite-Element Modeling and Experimental Verication of Spray-Cooling Process in Steel

R. THOMAS, M. GANESA-PILLAI, P.B. ASWATH, K.L. LAWRENCE and A. HAJI-SHEIKH An atomizer is a helpful tool that can be used to tailor the cooling rate of steel from the processing temperature in order to get desired properties. It is important to determine the temperature distribution in a specimen subjected to cooling by an atomized spray. A nite-element model for transient heat transfer and thermal-stress analysis is developed to determine the temperature and thermal-stress distribution. The results of the nite-element heat-transfer model are compared with a nite-difference model. The heat-transfer model describes the heat-transfer processes in an AISI 4140 steel cylinder subjected to controlled atomized spray cooling from an initial temperature of 1273 K. The temperature elds predicted by the model are used both to predict the resulting microstructure using continuous cooling transformation (CCT) diagrams and as an input for the thermal-stress model to predict the occurrence of quench cracks. The thermal-stress model incorporates temperature-dependent material properties, heat generation due to phase changes, elastoplastic behavior of steel, and the volumetric expansion associated with the formation of martensite. The results of the nite-element model are veried experimentally by recording temperature proles, obtaining micrographs, and recording the occurrence of quench cracks.

I.

INTRODUCTION

THE need for new materials has spurred innovation and research in metallurgical processing operations. In order to improve and optimize productivity and quality, industries such as forging, welding, casting, and heat treatment are adopting process-modeling techniques to reduce defects, scrap, design lead time, and cost.[1,2] In most forge shops, the forged parts are piled together and allowed to cool in air. This process is very slow and a large number of parts are produced before the rst part is cooled down and taken for preliminary inspection. Moreover, the nal product needs to be heat treated to meet the customers specication. Therefore, if specic properties can be engineered by microstructural control of alloys during the forming process itself, it would lead to substantial economic savings. Accelerated cooling by quenching has been demonstrated as a possible method to get desired properties.[36] However, this method is limited to low-carbon and microalloyed steels because of the possibility of quench cracks in medium- and high-carbon and alloy steels. Microalloyed steels, by virtue of their low alloy content, are not very hardenable and do not pose a signicant challenge. However, low-alloy steels contain larger amounts of alloying additions like Mo and Cr, which increase the hardenability of the steel and therefore increase the chances of quenchcrack formation. Furthermore, the quenching process does not facilitate the tailoring of cooling rates unless one uses a variety of quenching media and agitation.[7,8] Water spray cooling using jets of water was found useful
R. THOMAS, Research and Development Engineer, is with Komag, San Jose, CA 95131. M. GANESA-PILLAI, Graduate Student, P.B. ASWATH, Associate Professor, and K.L. LAWRENCE and A. HAJISHEIKH, Professors, are with the Department of Mechanical and Aerospace Engineering, University of Texas at Arlington, Arlington, TX 76019. Manuscript submitted March 23, 1995.
METALLURGICAL AND MATERIALS TRANSACTIONS A

in controlling the cooling rate of hot-rolled medium-carbon steel bars.[9] Accelerated cooling experiments in highstrength low-alloy (HSLA) steels using fog-jet nozzles indicate the possibilities of controlling the temperature gradient while cooling.[10] In an earlier work,[11] it was experimentally shown that an atomizer with a mixture of air and water could be successfully employed to tailor the hardness and tensile strength of a microalloyed steel. The relationships between composition, microstructure, and mechanical properties have been studied by previous investigators.[1215] Campbell et al.[13,14,15] have developed a mathematical model that incorporates the heat-transfer, transformation-kinetics, and property-structure-composition relationships to predict properties of plain carbon steels. For the analysis therein, the lumped heat-capacity method and the one-dimensional (1-D) implicit nite-difference method gave comparable results. The lumped heat-capacity model assumes that the temperature gradient between the surface and the interior is negligible, which is indeed true in air cooling, but the model is valid only when the Biot number (Bi) is small (Bi 0.1). This article deals with experiments and modeling of spray cooling of AISI 4140, which is a typical alloy steel, from high temperatures by a two-component (air-water) spray. The experimental procedure is discussed briey in Section II. Spray cooling is a new entrant in high-heat ux cooling and the heat-transfer relations are not well established as they are in single-phase ow. The atomized spray consists of small liquid droplets in a conical jet of air. By increasing or decreasing the amount of liquid in the spray, one can increase or decrease the cooling rate at the surface. Therefore, the cooling rate can range from that of forcedconvection air cooling to a quenching process. In metallurgical applications, water is the liquid of choice because of its low cost. The Biot number in such a process is very high, which means that there is a large gradient in temperature from the surface to the interior. For example, here the minimum conductivity (at the highest temperature)
VOLUME 29A, MAY 19981485

Table I. C 0.38 to 0.43 pct Cr 0.8 to 1.1 pct Mn

Nominal Composition of AISI 4140 Mo 0.15 to 0.25 pct P (Max) 0.035 pct S (Max) 0.04 pct Si 0.15 to 0.30 pct

0.75 to 1.0 pct

Fig. 1Schematic of the atomized spray-cooling setup used to cool the cylindrical specimens.

dict the occurrence of quench cracks. The data from the nite-difference heat-transfer analysis and the nite-element predictions of temperatures are compared to measured temperatures at a few selected sites after testing. This comparison is for verifying the nite-element model. Samples are removed from different locations of each test material, polished, and used for metallographic studies. The tests are performed to show microstructures and possible formation of cracks. A correlation is made between the predicted hardness of the steel using continuous cooling transformation (CCT) diagrams available in the literature and the hardness measured by Jominy quench tests at various locations. In addition, a correlation is made between the actual microstructure and the microstructure prediction based on the cooling rates. The results and conclusions are presented in Sections V and VI, respectively. II. EXPERIMENTAL APPARATUS AND PROCEDURE

is 45 W/m/K, the radius is 0.0191 m, and the heat-transfer coefcient is of the order of 105 W/m2, which makes Bi 0.1. Therefore, the lumped parameter model (valid only when Bi 0.1) is inappropriate and the heat-conduction equation needs to be solved. The heat-transfer coefcients at the surface have to be prescribed as boundary conditions. The relations that describe the surface heattransfer coefcient during various phases of the spray-cooling process are presented in Section III. These relations were obtained from experimental data and Monte Carlo inverse heat-conduction analysis by Buckingham and HajiSheikh.[16,17] The internal temperatures are computed by the explicit nite-difference formulation of the conduction equation in the cylindrical coordinate system. The article also demonstrates the prediction of quench cracks with commercial nite-element analysis (FEA) packages using available information on a complex heat-transfer phenomenon like spray cooling. The FEA method is currently the most popular modeling technique, because of the exibility in the denition of geometry, loading and boundary conditions. The tasks of building or updating a model and analyzing the results have now been greatly simplied by powerful FEA packages. The FEA package used in this work is ABAQUS, which provides an extensive library of functions to dene the complex phenomena in spray cooling.[18] The ABAQUS package allows detailed denition of the material, such as temperature-dependent properties, phase-transformation effects, and spatial and temperature variation of heat-transfer coefcients, which are required to describe the spray-cooling process. It is also very easy to incorporate new information as it becomes available in the future, in spray cooling or in phase-transformation kinetics, so that the predictions can be improved. The nite-element modeling described in Section IV of this article consists of two steps. The rst step describes the heat-transfer behavior of the material during spray cooling by determining the transient temperature distribution within the material. The temperature proles generated are used to predict the resulting microstructure of the material. Subsequently, these temperatures are used to calculate thermal stresses during cooling, which are used in turn to pre1486VOLUME 29A, MAY 1998

The material used for all experiments is AISI 4140 and its typical composition is shown in Table I. The experimental runs were conducted on an AISI 4140 steel bar, 0.0381 m in diameter and 0.1016-m long. A schematic of the spray-cooling apparatus is shown in Figure 1. The apparatus consists of a BETE 1/4XA-PR150-B atomizer, obtained from Bete Fog Nozzle Inc. (Greeneld, MA), that internally mixes compressed air and water in an annular region within the nozzle to form a ne conical spray. The water is supplied by a water pump and its ow rate is controlled by a needle valve. The ow rate is measured by a rotameter to aid manual ow control by the operator and by a ow meter, which sends the readings to a data acquisition system (DAS). The water pressure is monitored by a pressure transducer, which sends data to the DAS, and also by a pressure gage. Compressed air is supplied by an air compressor and is controlled by a ball valve. The air-ow rate is monitored, in the same manner as the water, by a ow meter, a pressure transducer, and a pressure gage. The air-ow meter provides an electronic display of the ow rate and also sends these readings to the DAS. The water and compressed-air lines meet at the atomizer, which uses the energy of the compressed air to atomize the water, and the mixture is expelled from the nozzle as a ne spray. More information on the nozzle characteristics is available in a previous publication.[11] The specimen is rst heated to a temperature of 1273 K in a furnace and then quickly transferred into a xture and exposed to the spray. During cooling, the specimen is located 0.41 m from the tip of the nozzle and in such a manner that it is upright and centered with respect to the spray. The particular distance was chosen because the spray was characterized only at that distance and the spray was fully developed with a fairly uniform distribution of droplets over a radial distance of 5 cm.[11] In actual practice, the spray completely engulfs the specimen. Specimens are cooled at four different cooling rates by varying the ow
METALLURGICAL AND MATERIALS TRANSACTIONS A

Table II. Specimen Number 1 2 3 4 5 6 7

Experimental Variables Water Flow Rate (kg/s) as received 0.00842 0.00602 0.00405 0.00184 0.00112 0

Air Flow Rate (kg/s) as received 0.000732 0.000698 0.000778 0.000819 0.000909 0.000997

11.5 8.6 5.13 2.24 1.1 0

during cooling. To withstand the high temperatures, thermocouples sheathed in steel are used. The thermocouples are xed to the locations using a high-temperature PYROPUTTY,* which is a mixture of steel powder and ceramic
*PYRO-PUTTY is a trademark of Aremco Products, Ossining, NY.

compounds. Following the spray-cooling runs, the specimens are cut for microstructural studies, and quench cracks, if any, are recorded. In addition, Rockwell C hardness measurements are made in a direction perpendicular to the spray and parallel to the spray at the midsection of the cylindrical specimen. III. HEAT-TRANSFER ISSUES IN SPRAY COOLING

Fig. 2Location of the thermocouples and coordinate conventions used in the heat-transfer and nite-element models. Hardness was measured along the lines shown in the pictorial view, in the plane of the thermocouple locations.

Fig. 3Schematic of the heat-transfer prole showing different heattransfer regions during spray cooling using an atomizer (used to derive Eq. [6]).

rates of water and air. The ow rates and pressures recorded for the four cooling rates are shown in Table II. The cooling rate increases with the increasing water-ow rate. A parameter , dened as the ratio of the water-mass-ow rate to the air-mass-ow rate, is used to represent each cooling rate. This parameter empirically relates the heat-transfer coefcients with the air- and water-ow rates. Four K-type thermocouples located as shown in Figure 2 are used to record the temperatures at these locations
METALLURGICAL AND MATERIALS TRANSACTIONS A

The heat-transfer process in spray cooling is complex and will not readily submit to numerical simulation. The factors that inuence the local heat-transfer rate are droplet distribution, thermal radiation, thermophysical properties, and kinematic quantities. The distribution of droplet size depends on air pressure, water-ow rate, and nozzle characteristics. The velocity eld in the spray has axial and radial dependence and is usually accompanied by plume oscillations. The thermophysical properties of the gas phase depend on the temperature and mole fraction of the constituents. The high-temperature nature of the heat-treating process introduces additional factors that can inuence the prediction of temperature. The determination of temperature- and location-dependent heat-transfer coefcients for the spray-cooling process is based on the results of Buckingham and Haji-Sheikh.[16,17] This section outlines the derivation of these coefcients for the specimen geometry under consideration. The heat-transfer process during spray cooling can be divided into three distinct regions: (1) a radiation-dominated region where, depending on the amount of water in the spray and the temperature of the surface, the water droplets will partially or totally vaporize before getting near the surface of the specimen and the air-water vapor-air mixture will cool the surface by convection; (2) a convectiondominated region, where, as the surface temperature decreases, the water droplets begin to arrive near the surface and enter the boundary layer over the specimen and the rate of heat transfer will increase rapidly; and (3) a transition region, which links the previous two regions together. In the convection-dominated region, the heat-transfer mechanisms are more complex and the heat-transfer coefcient is very large. However, when the Biot number is large, large errors in the value of the Biot number will have a small inuence on the computed internal temperature. A schematic form of the variation of heat ux with temperature at the surface of a cylinder subject to spray cooling is shown in Figure 3. This schematic form adapted from the results reported in Reference 17 is used in Section IIIC to derive the heat-transfer relations for the transition region. Since only the trends are important in the derivation, the values of the heat-transfer coefcient in Figure 3 do not matter. Surface heat-ux proles of a cylinder subjected to spray cooling are reported in the literature.[17,19] The peak heat ux is of the order of 106 W/m2. The temperature Tmax is dened as the temperature at
VOLUME 29A, MAY 19981487

which the heat ux q at the stagnation point is maximum. The radiation-dominated heat-transfer process exists at temperatures far above Tmax. In the radiation-dominated region, the liquid droplets will evaporate and heat transfer will be attributable to a homogeneous mixture of air and water vapor; it is assumed that the mole fractions have no spatial dependence. The Tmax is also the temperature at which the convective region starts. The heat-transfer coefcient at the stagnation point (h0) is calculated rst, and then, using approximate spatial relationships, the heat-transfer coefcient at any point at that particular time instant can be calculated as a function of h0. It is assumed that the gas is fully mixed and that there is no liquid in the spray for all locations with 90 deg. An accurate prediction of local values of the heat-transfer coefcient is a difcult task because the spatial variations depend on the distribution of droplet size, partial evaporation of liquid by radiation, and kinematic quantities. For this reason, only an estimation of the spatial variation of the heat-transfer coefcient is possible. The calculation of h0 is outlined in the following sections for each of the heattransfer regions. In general, the local surface heat transfer is computed by Eq. [1].
4 q h (Tw Ta) (T 4 w T )

mula for single-phase ow over a cylinder. For the range 104 Rem 2 105, the following Eq. [3][20] holds: hR,0 km 0.333 [82.81 0.003014 (Rem Pr m ) D0 0.333 2 3.265 109 (Rem Pr m )]

[3]

where hR,0 is the heat-transfer coefcient at the stagnation point for the radiation-dominated region. The thermophysical properties of air are taken from Incropera and DeWitt[21] and the thermophysical properties of water vapor from Haar et al.[22] The method for obtaining the thermophysical properties for a mixture of low-pressure gases is in Bird et al.[23] B. Convective Region When the stagnation-point temperature T0 is below Tmax, the convective region begins, and the water droplets reach the surface of the specimen and boil away. In this region, the heat-transfer coefcient at the stagnation point hc,0 is given by Eq. [4].[17] hc,0 1.9 105 kair (Tw,0 Ta)0.84 0.75 D0 [4]

[1]

where is the emissivity of alloy steel and is the Stefan Boltzmann constant. The local heat-transfer coefcient is h hR in the radiation-dominated region, and h hc when the phase-change process takes place in the boundary layer. A. Radiative Region The computations begin by using the heat-transfer relation for cross ow over a cylinder when Tw Tmax. The Reynolds number is dened by Eq. [2]. Rem

where Tw,0 is the wall temperature at the stagnation point, Ta is the stream temperature under adiabatic conditions, is the mass-of-water to mass-of-air ratio, hc,0 is the heattransfer coefcient at the stagnation point, D0 is the diameter of the cylinder, and kair is the thermal conductivity of air computed at the mean lm temperature. C. Transition Region The transition region is a poorly dened region that links the convective and radiative heat-transfer regions. The heattransfer coefcient at the stagnation point h0 is approximated by a linking function dened as follows: h0 a b exp [ (Tw,0 Tmax)] [5] where a, b, and are constants to be determined using the following three conditions: h0 hmax when Tw,0 Tmax; h0 hR,0 when Tw,0 Tmax is large; and dq/dt 0 or dq/dT 0 when Tw,0 Tmax. The form of the curve given by Eq. [5] is based on the trend of the heat-ux variation with temperature (Figure 3). The rst and the third conditions are based on the existence of the maximum heat ux and hence the maximum heattransfer coefcient when Tw,0 Tmax. The second condition stipulates that when Tw,0 is much greater than Tmax, h0 should tend to hR,0. The exponential term disappears when Tw,0 Tmax and therefore the form satises the requirements. It should be noticed that in this case, h0 is a function of hR,0, dened using radiative heat-transfer conditions. The values of a, b, and satisfying the conditions are as follows: a hR,0, b hmax hR,0, and hmax [(hmax hR,0)(Tmax Ta)] Substituting these values in Eq. [5] we get the following:
METALLURGICAL AND MATERIALS TRANSACTIONS A

au0 (1 ) D0 Tsat m Ta

[2]

where a is the density of the mixture at the adiabatic spray temperature and m is the viscosity of the mixture calculated at the mean bulk temperature. Some explanation is due here for the special denition of the Reynolds number. In the radiation-dominated region, the assumption is that all the liquid in the spray will evaporate before it arrives at the cylinder. In that case, the Nusselt-number data must exhibit a behavior similar to cross ow of a single-phase uid over a cylinder, where the Nusselt number depends on the Reynolds number and the Prandtl number. The process is very complex, since the heat transfer is attributable to a mixture of air and superheated water vapor. The water-vapor production contributes to an increase in the volume ow rate of the gas and a consequent widening of the jet. The factor (1 ) in the denition of Reynolds number accounts for the vapor production in the air, assuming that the air and water vapor are fully mixed. The quantity Ta/Tsat includes the effect of thermal expansion of the gas before reaching the boundary layer. When radiation dominates, the heat transfer is to a gaseous mixture, and therefore the heat-transfer coefcient at the stagnation point,hR,0, can be calculated using the for1488VOLUME 29A, MAY 1998

radiative and convective heat transfer. The surface of the cylinder away from the spray is assumed to be in contact with a mixture of water vapor and air only. The convention for the spatial variables and Z is shown in Figure 2. It should be noted that when 90 deg, h0 may be calculated through the convective or the radiative relations, depending on the temperature; i.e., if Tw Tmax, h0 hR,0, and if Tw Tmax, h0 hc,0. However, for 90 deg, h0 is calculated using the radiative relation. The spatial relations with measured in radians are as follows: h h0 1 0.5 (1 B) 4C cos 2 cos (1 0.18 Z ) 1 4C 4C
2

[7]

Fig. 4Variation of the heat-transfer coefcient with the angular coordinate .[17]

where B 0.385 0.0426 0, C cos 0, 0 1.995 0.016 0, 0 when 4, and 0 4 when 4. The angular dependence given previously is based on the heat-transfer data of Buckingham and Haji-Sheikh.[17] A plot of the spatial variation of the heat-transfer coefcient according to Eq. [7] is shown in Figure 4. In the radiationdominated region, h0 approaches hR,0. Using the aforementioned formulas the heat-transfer coefcient at any point on the surface of the cylinder can be calculated, given the heattransfer coefcient at the stagnation point. IV. The Finite-Element Model

Fig. 5Geometry of the axisymmetric nite-element model used to determine the temperature distribution in spray cooling.

The FEA is divided into two phases: heat-transfer analysis and thermal-stress analysis. Three models are evaluated for the heat-transfer analysis: an axisymmetric model, a twodimensional (2-D) model, and a three-dimensional (3-D) model (Figures 5, 6, and 7, respectively). The axisymmetric model assumes that the heat-transfer coefcient is independent of the coordinate whereas the 2-D model assumes that the heat-transfer coefcient is independent of the Z coordinate. Physically, these assumptions mean that there is no heat ow in the azimuthal direction for the axisymmetric case and in the Z direction for the 2-D case. It is shown subsequently that these approximations do not induce signicant error in the solution. Hence, the thermalstress analysis employs only an axisymmetric model. The boundary conditions imposed for the three models are shown in Figure 8. A. Heat-Transfer Module of the Finite-Element Model

Fig. 6Schematic of the 2-D nite-element model used to determine the temperature distribution.

h0 hR,0 (hmax h0) e where hmax FACT hmax hR,0

FACT

TT

w,0

Tmax Ta

max

[6]

The heat-transfer model denes the process of spray cooling and the thermal behavior of the material over the temperature range under consideration. The governing equation for heat transfer in the cylinder for all the three cases analyzed is as follows: kT q "' Cp T t [8]

D. Spatial Relationships Empirical spatial relations relate the heat-transfer coefcient at any point to that at the stagnation point, h0. The surface of the cylinder facing the spray experiences both
METALLURGICAL AND MATERIALS TRANSACTIONS A

The initial condition for all cases is as follows: at t 0, T Tin 1273 K. The term q is the heat generated per unit volume. This term is invoked when there is an exothermic or endothermic phase transformation. For the axisymmetric case, the following holds:
VOLUME 29A, MAY 19981489

kT T r T k r T k z T k z k

1 T kr r r r

z k T z
[9]

h (Tw Ta) 0 0 0

at r 0.01905 m, t 0 at r 0, t 0 at z 0, t 0 at z 0.0508 m, t 0

For the 2-D case, the following holds: kT T r T k r T k T k k 1 T kr r r r

T r1 k
2

h (Tw Ta) 0 0 0

at r 0.01905 m, t 0 at r 0, t 0 at 0, t 0 at , t 0 [10]
Fig. 7Schematic of the 3-D nite-element model used to determine the temperature distribution.

For the 3-D case, the following holds:


kT 1 T kr r r r

T r1 k z k T z
2

T z T k r T k r T k k

at z 0 and z 0.0508 m, t 0 at r 0.01905 m, t 0 [11]


Fig. 8Boundary conditions used in the nite-element models.

h (Tw Ta) 0 0

at r 0, t 0 at 0 and , t 0

The thermal behavior of materials is dened by specifying the temperature dependence of thermal conductivity and specic-heat capacity[24] and the latent heat effects due to phase transformations (Figure 9). The heat generated by the austenite-to-ferrite/pearlite phase transformation results in a sharp peak for the specic-heat curve. The quantity of heat generated is specied by calculating the area under the peak of the dotted line in the Cp vs T curve. This latent heat effect is suppressed when the cooling rate is too high for the transformation to take place and hence is incorporated in the model only for the lower cooling rates. In the cases where the phase transformation does occur, a latent heat of 106 J/kg is dened between 1013 and 1053 K. The region within which the heat generation is specied is obtained by determining the temperature range in which the inection occurs in the experimental cooling curve.[24] The spray-cooling process is characterized by specifying location- and temperature-dependent heat-transfer coefcients, as dened previously. This denition is straightforward for the 3-D case because both the spatial coordinates
1490VOLUME 29A, MAY 1998

Fig. 9Temperature dependence of the thermal conductivity and specic heat of AISI 4140.[24]

are explicitly dened in the model geometry. For the 2-D case, the heat-transfer coefcients are dened by specifying that Z 0 in the spatial relationships. This is justied because the heat-transfer coefcient varies by less than 10 pct in the range of Z considered (Figure 4.). The variation of
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 10Temperature dependence of the tensile strength and elastic modulus of AISI 4140 steel.[24]

temperatures considered in this article. In the calculations, the elastic modulus and tensile strength are considered to be the same as those at 800 K, for all higher temperatures. This is shown to be a conservative estimate. The elastic modulus decreases with increasing temperature. The tensile stress is calculated by the product of the elastic modulus and the thermal strain. If the calculated stress is more than the ultimate tensile strength, the model predicts a crack. Since an elastic modulus higher than the true value is used, the stress is being overestimated. It is true that the tensile strength to which the calculated stress is compared is also higher than the true value. However, tensile strength at 860 K is already very small and keeping it the same at higher temperatures will not alter the predictions at all. Therefore, the predictions are conservative and satisfactory. Phasetransformation effects are modeled by appropriately modifying the temperature dependence of the thermal-expansion coefcients, as explained in the following sections. Martensitic transformation The 6 pct volumetric expansion accompanied by the martensitic transformation has a signicant effect on the thermal-stress distribution. The reaction takes place at 615 K in AISI 4140 when the cooling rate is high. At temperatures higher than 615 K, the thermal-expansion coefcients are as shown by the solid line in Figure 11.[24] From Eq. [12], it can be seen that a 6 pct volumetric expansion is equivalent to a 2 pct linear expansion.
3 L3 V (Li L)3 Li3 3 L f Li 6 pct V L3 L3 L i i

[12]

Considering a 1-D case of expansion, the thermal-expansion coefcient is related to the initial and nal lengths by Eq. [13] as follows: L f0 Li (1 0T )
Fig. 11Thermal expansion coefcient with and without phase transformation in AISI 4140 steel.
f 0 i

[13]

heat-transfer coefcient with is signicant and hence choosing an arbitrary value for would lead to large errors. Therefore, for the axisymmetric case, the average of the heat-transfer coefcients over the range 0 is calculated by integrating Eq. [7] from 0 to and dividing by , and this average value is specied over the circumference. The temperature proles at different locations for these experiments, calculated by the nite-difference method, are reported in an earlier publication.[19] B. The Thermal-Stress Module This module calculates the thermal-stress distribution within the specimen during cooling. The model imports the temperatures from the heat-transfer analysis to calculate stresses induced due to thermal gradients and phase transformations, which in turn are used to predict the occurrence of quench cracks in the specimen. This analysis is conducted using only the axisymmetric model, as mentioned previously. Elastoplastic-material behavior is specied by dening temperature-dependent material properties such as elastic modulus and tensile strength (Figure 10). Information on these properties is not available for the range of
METALLURGICAL AND MATERIALS TRANSACTIONS A

where L is the nal length without transformation, L is the initial length, 0 is the thermal-expansion coefcient without considering transformation, and T is the change in temperature from the initial condition. In Eq. [13], 0 refers to the thermal-expansion coefcients given by the solid line in Figure 11, which do not consider the effect of expansion due to the martensitic transformation. For temperatures below 615 K, 0 values are modied to account for the expansion in the following way. If LfT represents the nal length, taking into account the transformation, LfT is related to Lf0 by Eq. [14].
f f LT L0 1.02

[14]

This arises because of the 2 pct linear expansion due to the martensitic transformation. Hence, the following hold
f L fT L 0 1.02 Li (1 T T )

[15]

where T is the thermal-expansion coefcient considering the expansion due to the martensitic transformation, which gives the following: Li (1 0T ) 1.02 Li (1 T T ) Simplifying, we get the following: 0.02 1.02 0T T [16]

[17]

VOLUME 29A, MAY 19981491

This relation is used to calculate the modied thermalexpansion coefcients, and these values are compared to the original thermal-expansion coefcients in Figure 11. V. RESULTS AND DISCUSSION

The heat-transfer module of ABAQUS predicts the temperature distribution within the specimen as a function of time at the various cooling rates. These temperature proles are used to predict the resulting microstructure at the location under consideration and serve as an input for the thermal-stress module. The predictions made by FEA are compared with experimental results in the following sections. A. Temperature Proles 1. Finite-difference heat-transfer-analysis predictions The temperatures at various locations were calculated using the explicit nite-difference scheme and compared with the experimental data. The temperature proles and surface heat-ux history are reported in detail by Thomas et al.[19] A cooling curve at location 2 (the center of the cylinder) is shown in Figure 12 for comparison and discussion. Despite the complications involved in modeling the heat transfer in an atomized spray, the predictions match the experiments reasonably well. The cooling rate, as expected, is highest for 11.5, which has the largest water-ow rate. At 11.5 there is a direct transformation of austenite to martensite and bainite, and hence there is no bump in the cooling curve. However, for 2.24 and 1.1, the match between the experiments and the heat-transfer models is not accurate at temperatures below 750 K. In Figure 12, the experimental temperature proles at the lower cooling rates have a tendency to level off at around 750 K. The cooling rate is slowed down here because at this temperature the formation of pearlite is favored, and the latent heat due to the austenite-to-pearlite phase transformation is released. The nite-difference heat-transfer model was not able to predict the temperature prole well, because the latent heat was not included in the calculations. The latent heat was not included because of the dependence of the transformation temperature on the cooling rate and chemical composition. At the higher cooling rates, there is a direct transformation from austenite to martensite and bainite. A latent heat corresponding to this transformation will be released, but because the cooling rates are very high, the latent heat is quickly transferred to the external ow. Therefore, there is no response to this heat generation in the temperature prole, whereas there is a clear slowing down of the cooling process for the lower cooling rates, wherein the formation of pearlite is favored. A sharp increase in cooling rate at approximately 615 K in the heat-transfer model is also seen in the experimental data. This is associated with the start of the convective region when droplets of water are able to arrive at the boundary layer without completely evaporating. 2. Finite-element predictions Temperature proles are generated by the axisymmetric, 2-D, and 3-D models and are compared with experimental results in Figures 13, 14, and 15, respectively. For the axisymmetric model shown in Figure 13, 2.24 and 1.1
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 12Comparison of experimental cooling curves with the predictions from the heat-transfer model for the thermocouple at location 2.

Fig. 13Comparison of experimental cooling curves with the predictions from the axisymmetric nite-element model for the thermocouple located at position 2.

Fig. 14Comparison of the experimental cooling curves with predictions from the 2-D nite-element model for the thermocouple located at location 2.
1492VOLUME 29A, MAY 1998

Fig. 15Comparison of the experimental cooling curves with predictions from the 3-D nite-element model for the thermocouple located at position 2.

Fig. 16Temperature history at location 2 calculated using the axisymmetric nite-element model for 1.1 and 2.24 superimposed on the CCT diagram for AISI 4140 steel.

have an inection in the cooling curve due to heat generated by the austenite-pearlite transformation. This latent heat effect is accurately reproduced by the model. At the higher cooling rates, i.e., for 5.13 and 11.5, the austenite-pearlite transformation does not occur and hence the associated inection is absent. The 2-D and 3-D model predictions are shown in Figures 14 and 15, respectively. The 2-D model assumes that there is no heat ux in the Z direction of the specimen, whereas the 3-D model does not make this assumption. Since the 2-D and 3-D models predict almost identical results, it can be concluded that there is negligible heat ux in the Z direction. There is no information available in the literature for the latent heat of formation for the phase transformation from austenite to martensite for AISI 4140 steel. Therefore, the nite-element calculations were initially carried out including a ctitious but large value for the latent heat of formation for the austenite-to-martensite transformation. There was no change in the temperature proles when compared with the calculations when the latent heat was not included. Therefore, in the subsequent nite-element calculations, the latent heat was included for the austenite-to-pearlite transformations observed at the lower cooling rates and neglected for the austenite-to-martensite transformation observed at higher cooling rates. It is observed that despite the large variation of the heattransfer coefcient with , the axisymmetric model gives the best match and is therefore used to predict the microstructures. The 3-D model does not give any specic advantage and therefore is not used in subsequent calculations. This is not surprising, considering the complexity of the heat-transfer phenomenon in spray cooling. B. Microstructures Once the cooling curves are known, the resulting microstructure can be predicted by superimposing these cooling curves on a CCT diagram of AISI 4140 steel.[25] When a cooling curve is superimposed on the CCT diagram, the phases that form can be determined by the regions that the curve passes through. The steel is initially made entirely of austenite (A). As the austenite cools it transforms to other
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 17Temperature history at location 2 calculated using the axisymmetric nite-element model for 5.13 and 11.5 superimposed on the CCT diagram for AISI 4140 steel.

phases, namely, pearlite (P), ferrite (F), bainite (B), and martensite (M). The labeled solid lines represent the start of the various transformations, namely, austenite-to-ferrite (Fs), austenite-to-pearlite (Ps), austenite-to-bainite (Bs), and austenite-to-martensite (Ms). The labeled regions show the phases that exist with it. The dashed lines represent the fraction of the total austenite that has transformed. Combining these two, it is possible to nd out which phases and how much of each phase will form at a particular cooling rate. 1. 1.1 At this cooling rate, the curve enters the region of the CCT diagram (Figure 16). The curve rst crosses Fs into the ferrite-forming region and then crosses Ps into the pearlite-forming region. At this point, only 1 pct of the austenite has transformed and hence the fraction of ferrite formed is still negligible. However, as discussed previously, ferrite continues to form along with pearlite between the Ps and Bs lines. When the curve crosses Bs into the bainitic region,
VOLUME 29A, MAY 19981493

(a)

(b)

(c)

(d)

Fig. 18Optical micrographs of the microstructure of the cylindrical specimens cooled at (a) 1.1 showing 90 pct pearlite and 10 pct bainite, (b) 2.24 showing 50 pct pearlite and 50 pct bainite, (c) 5.13 showing bainite and martensite, and (d ) 11.5 showing martensite and bainite.

Table III.

Summary of Microstructure Prediction Results Predicted Microstructure Based on Heat-Transfer Model PF 80 pct (F P) 20 pct B 50 pct (F P) 50 pct B 15 pct F 60 pct B 25 pct M 10 pct F 40 pct B 50 pct M 10 pct B 90 pct M 100 pct M

0 1.1 2.24 5.13 8.6 11.5 Quench

Experimentally Observed Microstructure PF 90 pct P 10 pct B 50 pct P 50 pct B BM BM BM 100 pct M

about 80 pct of the austenite has transformed into ferrite pearlite. The remaining austenite then transforms to bainite. The predicted microstructure is therefore 80 pct ferrite pearlite and 20 pct bainite. The actual microstructure is shown in Figure 18(a). The microstructure is predominantly dark-etched ne-grained pearlite with a few light-etched patches of bainite. The characteristic needles of bainite can be seen within the light etching. 2. 2.24 Figure 16 shows the curve rst crossing Fs, and then, after 1 pct transformation, across Ps. After 50 pct of the
1494VOLUME 29A, MAY 1998

austenite has transformed into ferrite pearlite, the remaining austenite transforms to bainite. The predicted microstructure is therefore 50 pct ferrite pearlite and 50 pct bainite. The observed microstructure (Figure 18(b)) conrms these predictions very clearly. The dark etching shows very ne pearlite and the light etching shows the needlelike bainitic structure. The proportion of each phase is approximately 50 pct. 3. 5.13 As shown in Figure 17, after about 15 pct of the austenite transforms to ferrite, bainite starts to form. When the curve
METALLURGICAL AND MATERIALS TRANSACTIONS A

C. Thermal Stresses The thermal-stress model generates the stress distribution within the material during cooling. These stress distributions, along with the maximum-stress failure criterion, are used to predict the occurrence of quench cracks. The maximum principal stresses are plotted against their associated temperatures and are shown in Figures 20 and 21 for 5.13 and 11.5, respectively. These maximum stresses are compared with the tensile strength of AISI 4140, which is also a function of temperature.[24] The maximum-stress criterion states that if any of the three principal stresses in the material exceed the tensile strength, failure will occur. Since the thermal stresses at lower cooling rates are lower and no quench cracks are predicted, the stress distributions are not shown for 1.1 and 2.24. The effect of the martensitic transformation on thermalstress and radial-crack formation is shown schematically in Figure 19. When the material cools at a high cooling rate, there is a large thermal gradient between the core and the surface of the cylinder. The surface reaches the martensitic transformation temperature rst and forms martensite. Martensite is a hard and brittle phase and its formation is accompanied by a 6 pct volumetric expansion. When the surface transforms to martensite, the resulting expansion is not resisted because the core is at a higher temperature and therefore is soft. At a later time, the core of the cylinder starts to form martensite. In this case, the expansion of the core is resisted by the hard surface, which is already martensite. This resistance by the surface induces a tensile hoop stress on the surface that can cause a radial quench crack to form. It is important to note that the formation of martensite must be accompanied by large thermal gradients between the core and the surface for the transformation to induce thermal stresses. If martensite forms all over the cylinder at the same time, the thermal stresses that are induced will be much smaller. The steel must also possess sufcient hardenability to form martensite at the surface as well as the core, and AISI 4140 does have high hardenability. 1. 5.13 Figure 20 shows the variation of maximum stress during cooling, and it is compared with the tensile strength. In Table III, it is shown that martensite does form at this cooling rate. The thermal-expansion coefcients are therefore appropriately modied for this model. However, this does not have any effect on the maximum stresses shown in the gure, which indicate that the cooling rate for this case is not high enough for the martensitic transformation to have an effect on the thermal-stress distribution. The hoop stress and longitudinal stress are comparable and larger than the radial stress. Because the maximum stresses do not approach the tensile strength at any point, the model predicts that the material will not fail. No quench cracks were observed during experiments at this cooling rate. 2. 11.5 In Figure 21, the maximum stress is plotted against the temperature at the location where the maximum stress occurred. This plot shows the effect of the martensitic transformation on thermal-stress distribution. Around the temperature at which martensite forms, which is 615 K, all three maximum stresses shoot up to very close to the tensile
VOLUME 29A, MAY 19981495

Fig. 19Schematic diagram of the formation of radial cracks due to martensitic transformation.

Fig. 20Maximum thermal stress and ultimate tensile stress calculated using the axisymmetric nite element against the corresponding temperature during cooling for 5.13.

crosses Ms, 75 pct of the austenite has transformed, which means that 60 pct of the transformed product is bainite. The remaining austenite transforms to martensite and hence the nal predicted microstructure is 15 pct ferrite, 60 pct bainite, and 25 pct martensite. The observed microstructure (Figure 18(c)) is almost entirely needle like, being a combination of bainite and martensite. Here, it is difcult to differentiate between the martensitic needles and the bainitic needles and hence no estimates on percentages are made. 4. 11.5 Figure 17 shows that after 1 pct of the austenite transforms into ferrite, bainite begins to form. After 10 pct of the austenite has transformed into bainite, the remaining austenite forms martensite. The predicted nal microstructure is therefore 10 pct bainite and 90 pct martensite. The actual microstructure (Figure 18(d)) shows a predominantly martensitic microstructure. A summary of the microstructure predictions is shown in Table III.
METALLURGICAL AND MATERIALS TRANSACTIONS A

is conrmed by Figure 22, which is a picture of a typical radial crack that occurred in specimens at this cooling rate. D. Rockwell Hardness Hardness proles of the spray-cooled specimens, taken in the direction perpendicular to the spray, are shown in Figure 23. The hardness increases with from Rc 30 for the as-received material to about Rc 60 for water quenching. There is little variation in hardness between the surface and the core of the cylinder for all cooling rates. A jump in hardness is observed between 2.24 and 5.13. This is because, as shown in previous sections, martensite begins to form at 5.13, leading to an increase in hardness. Table IV lists the measured cooling rate and that predicted by the heat-transfer analysis at 973 K, and the hardness at room temperature. For a given cooling rate, the Rockwell C hardness can be found from a standard Jominy end-quench table for 4140 steels. Comparison of the data in Table IV indicates that there is good match between the measured hardness and the hardness determined from a knowledge of the cooling rates determined by the heattransfer models. Therefore, the nite-element model, along with the available charts, is a convenient method of determining the microstructure and hardness at any location within the specimen, without the need for measurement of temperature and cooling rate. VI. CONCLUSIONS

Fig. 21Maximum thermal stress and ultimate tensile stress calculated using the axisymmetric nite element against the corresponding temperature during cooling for 11.5.

The following statements summarize the analysis of the spray-cooling process and the scheme proposed to model the process. 1. The air/water atomizer can be used to tailor the cooling rate from the forging temperature in order to obtain desired properties in a one-step process from the forging operation without any post heat treatment. 2. The proposed heat-transfer model provides a reasonable estimate of the temperature prole in a AISI 4140 steel bar during cooling by an atomized spray of air and water. 3. The temperature distribution in spray-cooled steel can be predicted to reasonably good accuracy using niteelement modeling. The latent heat effects of the phase transformations and temperature dependence of material properties can be effectively incorporated in the model. 4. Resulting microstructures in spray-cooled steel can be predicted by means of the CCT diagram in combination with predicted temperature proles. 5. Hardness at any location in spray-cooled steel can be predicted by means of Jominy end-quench diagrams in combination with predicted temperature proles. 6. The thermal-stress model accurately predicts the occurrence of quench cracking. 7. Martensitic transformation and its effect on thermal stress can be incorporated. ACKNOWLEDGMENTS The authors gratefully acknowledge the support of the NASA/UTA Center for Hypersonic Research, The MAE
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 22Typical radial crack found in specimens for 11.5

Fig. 23Rockwell hardness proles along the diameter perpendicular to the direction of the spray for different values of .

strength. The maximum hoop stress is the largest of the three stresses and hence a radial crack is predicted in the material. This is consistent with the earlier discussion and
1496VOLUME 29A, MAY 1998

Table IV.

Measured and Predicted Cooling Rates at 973 K and Hardness at Room Temperature Predicted Cooling Rate at 973 K (K/s) 7 to 14 1.52 to 1.91 1.07 to 1.13 Predicted Hardness from Literature, Rc (Based on Experimentally Measured Cooling Rate) 47 to 53 38 38 Predicted Hardness from Literature, Rc (Based on Finite Difference Analysis) 46 to 50 38 38 Measured Hardness (Rc) 53 to 56 41 35

11.5 2.24 0

Experimental Cooling Rate at 973 K (K/s) 8 to 20 2.1 to 2.3 2.1 to 2.3

Department, and Trinity Forge Inc., who supplied the materials for this work.

surface condition ambient temperature

a, b Bi D0 h kair ks Cp mair mwater Nu Re Pr qs Q r R t T Ti Tsat u0 V L Li Lf0 LfT

NOMENCLATURE constants in Eq. (5) Biot number, hR/ks diameter of the cylinder, 2R surface heat-transfer coefcient, W/(m2K) thermal conductivity of air, W/(mK) thermal conductivity of steel, W/(mK) specic heat capacity, J/(kgK) Air, mass-ow rate, kg/s Water, mass-ow rate, kg/s Nusselt number, hD0/ka Reynolds number, [au0(1 )D0/m](Tsat/Ta) Prandtl number, Cp/k surface heat ux, W/m2 volume ow rate, L/s radial coordinate radius of cylinder, m time, s temperature, K initial temperature, K saturation temperature, K maximum velocity 0.41 m from nozzle exit, m/s volume, m3 length, m initial length, m nal length without martensitic transformation, m nal length incorporating the effects of martensitic transformation, m

Microstructure A austenite B bainite austenite-to-bainite start temperature Bs F ferrite austenite-to-ferrite start temperature Fs M martensite austenite-to-martensite start temperature Ms P pearlite austenite-to-pearlite start temperature Ps

REFERENCES
1. S.C. Jain: J. Met., 1994, vol. 46 (9), p. 16. 2. B.A. Mueller and P. Monaghan: J. Met., 1994, vol. 46 (9), pp. 1720. 3. K. Amano, T. Hatomura, C. Shiga, T. Enami, and T. Tanaka: Proc. Int. Symp. on Accelerated Cooling of Rolled Steel, Winnipeg, Canada, 1987, G.E. Ruddle and E.F. Crawley, eds., Proc. Metallurgical Society of the Canadian Institute of Mining and Metallurgy, Pergamon Press, Elmsford, NY, 1988, vol. 3, pp. 43-56. 4. R.F. Price and A.J. Fletcher: Met. Technol., 1980, vol. 7 (5), pp. 203-11. 5. E.A. Almond, P.S. Mitchell, and R.S. Irani: Met. Technol., 1979, vol. 6 (6), pp. 205-14. 6. M.F. Mekkawy, K.A. El-Fawakhry, M.L. Mishreky, and M.M. Eissa: Mater. Sci. Technol., 1990, vol. 6 (1), pp. 28-36. 7. K.-E. Thelning: J. Heat Treating, 1983, vol. 3 (2), pp. 100-07. 8. C.E. Bates: J. Heat Treating, 1988, vol. 6 (1), pp. 27-45. 9. T. Ohshiro, Y. Maeda, S. Kuchiishi, T. Ikeda, and H. Sawada: Proc. Int. Symp. on Accelerated Cooling of Rolled Steel, Winnipeg, Canada, 1987, G.E. Ruddle and E.F. Crawley, eds., Proc. Metallurgical Society of the Canadian Institute of Mining and Metallurgy, Pergamon Press, Elmsford, NY, 1988, vol. 3, pp. 283-99. 10. M. Fegredo: Met. Technol., 1977, vol. 4 (9), pp. 417-24. 11. S.R. Pejavar and P.B. Aswath: J. Mater. Eng. Performance, 1994, vol. 3 (2), pp. 234-247. 12. A.J. DeArdo, J.M. Gray, L. Meyer, and S.C. Jain: Niobium 81, Proc. Int. Symp. Niobium 81, San Francisco, 1981, H. Stuart, ed., 1984, The Metallurgical Society of the AIME, Warrendale, PA, p. 685. 13. P.C. Campbell, E.B. Hawbolt, and J.K. Brimacombe: Metall. Mater. Trans. A, 1991, vol. 22A, pp. 2769-78. 14. P.C. Campbell, E.B. Hawbolt, and J.K. Brimacombe: Metall. Trans. A, 1991, vol. 22A, pp. 2779-90. 15. P.C. Campbell, E.B. Hawbolt, and J.K. Brimacombe: Metall. Trans. A, 1991, vol. 22A, pp. 2791-2805. 16. F.P. Buckingham and A. Haji-Sheikh: ASME HTD-Vol. 259, 1993, pp. 33-41. 17. F.P. Buckingham and A. Haji-Sheikh: J. Heat Transfer, 1995, vol. 117, pp. 1018-28. 18. ABAQUS Users Manual, HKS Inc., Pawtucket, RI, 1985. 19. R. Thomas, M. Ganesa-Pillai, P.B. Aswath, and A. Haji-Sheikh: ASME HTD-Vol. 289, 1994, pp. 31-41. 20. W.H. Giedt: Trans. ASME, 1949, vol. 71, pp. 375-81. 21. F.P. Incropera and D.P. DeWitt: Fundamentals of Heat and Mass Transfer, John Wiley & Sons, Inc., New York, NY, 1990, p. A15. 22. L. Haar, J.S. Gallagher, and G.S. Kell: NBS/NRC Steam Tables: Thermodynamic and Transport Properties and Computer Programs
VOLUME 29A, MAY 19981497

Greek 0 Thermal-expansion coefcients of steel without martensitic transformation T thermal-expansion coefcients of steel with martensitic transformation emissivity of cylinder angle in cylindrical coordinates viscosity coefcient, Ns/m2 w density of water, kg/m3 a density of mixture at adiabatic spray temperature, kg/m3 StefanBoltzmann constant mass-ow-rate ratio, mwater/mair Subscripts a gas-stream temperature c convection-dominated region m mixture max at maximum heat ux 0 stagnation point R radiation-dominated region
METALLURGICAL AND MATERIALS TRANSACTIONS A

for Vapor and Liquid States of Water in SI Units, Hemisphere Publishing Corporation, Washington, DC, 1984, pp. 261-70. 23. R.B. Bird, W.E. Stewart, and E.N. Lightfoot: Transport Phenomena, John Wiley & Sons, Inc., New York, NY, 1960.

24. High Temperature Property Data: Ferrous Alloys, M.F. Rothman, ed., ASM INTERNATIONAL, Metals Park, OH, 1988, pp. 3-10-14. 25. Atlas of Isothermal and Cooling Transformation Diagrams, ASM, Metals Park, OH, 1977, p. 391.

1498VOLUME 29A, MAY 1998

METALLURGICAL AND MATERIALS TRANSACTIONS A

Vous aimerez peut-être aussi