Sie sind auf Seite 1von 9

6th World Congresses of Structural and Multidisciplinary Optimization Rio de Janeiro, 30 May - 03 June 2005, Brazil

Evaluation of heat transfer coefficients during upward and downward transient directional solidification of Al-Si alloys
Jos Eduardo Spinelli 1 Ivaldo Leo Ferreira 1 Amauri Garcia 1
(1) Department of Materials Engineering, State University of Campinas - UNICAMP PO Box 6122 - 13083970 - Campinas, SP, Brazil *Email address: amaurig@fem.unicamp.br Aluminum alloys with silicon as a major alloying element, consist a class of alloys, which provides the most significant part of all shaped castings manufactured. This is mainly due to the outstanding effect of silicon in the improvement of casting characteristics, combined with other physical properties such as mechanical properties and corrosion resistance. In general, an optimum range of silicon content can be assigned to casting processes. For slow cooling rate processes (sand, plaster, investment) the range is 5 to 7 wt%, for permanent molds 7 to 9% and for die castings 8 to 12%. Since most castings parts are produced considering no dominant heat flow direction during solidification, it seems to be adequate to examine both upward and downward growth directions in order to better understand foundry systems. The way the heat flows across the metal/mold interface strongly affects the evaluation of solidification, and plays a remarkable role in the structural integrity of castings. Gravity or pressure die casting, continuous casting and squeeze casting are some of the processes where product quality is more directly affected by the interfacial heat transfer conditions. Once information in this area is accurate, foundrymen can effectively optimize the design of their chilling systems to produce sound castings. The present work focuses on the determination and evaluation of transient heat transfer coefficients from the experimental cooling curves during solidification of Al 5, 7 and 9 wt % Si alloys. The method used is based on comparisons between experimental data and theoretical temperature profiles furnished by a numerical solidification model, which applies finite volume techniques. In other words, the resulting data were compared with a solution for the inverse heat conduction problem. The necessary solidification thermodynamic input data were obtained by coupling the software ThermoCalc Fortran interface with the solidification model. A comparison between upward and downward transient metal/mold heat transfer coefficients is conducted. Keywords: Inverse heat flow problem, metal/mold heat transfer coefficient, mathematical modeling, directional solidification, Al-Si alloys. 1. Introduction The computer simulation of freezing patterns in castings has done much to broaden our understanding of casting and mold system design. The structural integrity of shaped castings is closely related to the time-temperature history during solidification, and the use of casting simulation could do much to increase this knowledge in the foundry industry. However, some uncertainties must be eradicated, particularly heat transfer at the metal/mold interface. The accurate knowledge of interfacial heat transfer coefficients is necessary for accurate modeling of castings. The ability of heat to flow across the casting and through the interface from the casting to the mold directly affects the evolution of solidification and plays a notable role in determining the freezing conditions within the casting, mainly in foundry systems of high thermal diffusivity such as chill castings. Gravity or pressure die castings, continuous casting and squeeze castings are some of the processes where product soundness is more directly affected by heat transfer at the metal/mold interface. Several studies have attempted to quantify the transient interfacial metal/mold heat transfer coefficient, hi, emphasizing the different factors affecting heat flow across such interface during solidification [1-6]. These factors include the thermophysical properties of the contacting materials, the casting geometry, the orientation of the casting-mold interface with respect to gravity (contact pressure), mold temperature, pouring temperature, the roughness of mold contacting surface, mold coatings, etc [3-4, 6-9]. Many relevant heat transfer mechanisms are associated with a relative movement of the casting and mold surfaces (initially in contact), which can cause, for instance, the appearance of a continuous gap. As a result of that imperfect junction between the two surfaces a temperature drop is created. Consequently, a less efficient condition of heat extraction is achieved. Figure 1 shows a schematic representation of two contacting surfaces. The heat flow across the casting-mold interface can be characterized by a macroscopic average metal-mold interfacial heat transfer coefficient (hi), given by: hi = q A( TIC TIM ) (1)

Figure 1. Representation of heat flow across mold-casting interface, where: q (W) is the average heat flux across the interface; TIC and TIM are, respectively, casting and mold surface temperatures (K). Most of the methods of calculation of time-dependent hi existing in the literature are based on numerical techniques generally known as methods of solving the Inverse Heat Conduction Problem, IHCP [10-12]. The IHCP method is based on a complete mathematical description of the physics of the process, supplemented with experimentally obtained temperature measurements in metal and/or mold. The inverse problem is solved by adjusting parameters in the mathematical description to minimize the difference between the model-computed values and the experimental measurements. The purpose of this study is to investigate the influence of direction of solidification on the time-varying interfacial heat transfer coefficient, hi. Downward and upward directional solidification systems were designed so that experiments could be carriedout with Al-Si alloys. Besides, the experimental temperatures were compared with simulations furnished by a numerical model for the determination of transient hi profiles. 2. Numerical modeling

Governing Equations
The numerical model used to simulate inverse segregation profiles is based on the previously proposed by Voller [13]. Modifications to this numerical approach have been incorporated to allow the use of different thermophysical properties for the liquid and solid phases, as well as the mushy zone (it can deal with temperature and concentration dependent thermophysical properties), to treat variable metal/mold interface heat transfer coefficient and to account a space dependent initial melt temperature profile. A time variable metal/mold interface heat transfer coefficient introduces a non-linearity condition at the z = 0 boundary. In addition, a variable space grid is used to assure the accuracy of simulation results without considerably raising the number of spatial nodes, when comparing with experimental data. Considering the previous exposed, the vertical unidirectional solidification of binary hypoeutectic Al-5 wt% Si, Al-7 wt% Si and Al-9 wt% Si alloys is our target problem. At time t < 0 , the alloys are in the molten state at the nominal concentration C0 and initial temperature distribution T0 ( z ) = a.z 2 + b.z + c , contained in the insulated mold defined by 0 < z < Z b according to Figure 2. Solidification begins by cooling the melt from the chill ( z = 0 ) until the temperature drops bellow the eutectic temperature TE . At times liquid. t > 0 , three transient regions are formed: solid, solid+liquid (mushy zone) and

Figure 2. Schematic ingot initial melt temperature distribution for upward solidification (t = 0).

To develop a numerical solution for the equations of the coupled thermal and solutal fields, we made the following assumptions: i. The domain is one-dimensional, defined by 0 < z < Z b , where Z b is a point far removed from the chill; ii. The solid phase is stationary, i.e., once the solid has formed it has zero velocity; iii. Due to the relatively rapid nature of heat and mass diffusion in the liquid, within a representative elemental averaging volume, the liquid concentration C L , the temperature T, the liquid density L and the liquid velocity u L are constant [14]; iv. The partition coefficient k0 and liquidus slope mL, are obtained from the ThermoCalc software*; v. Equilibrium conditions exist at the solid/liquid interface, i.e., the temperature and concentrations fulfill the equations: (2) T = TF m L C L and C* = k 0 C L , (3) S where TF is the fusion temperature of the pure solvent [K] and C* is the solid concentration at the interface; S vi. Thermophysical properties such as: specific heats, c S and c L , thermal conductivities k S and k L , and densities S and

vii.

L , are constants within each phase, but discontinuous at the solid-liquid boundary. The latent heat of fusion is taken as the difference between phase enthalpies = H L H S ; The metal/mold thermal resistance varies with time, and is incorporated in a global heat transfer coefficient defined as hi [15].

(*)The ThermoCalc software [16] can be used to generate equilibrium diagrams and through ThermoCalc interface for Fortran or C++ it is possible to recall those data generated by the software in order to provide model accurate results. Using the above assumptions, the mixture equations for binary solidification read: Energy
g cT + ( L c L uT ) = ( kT ) S T t

(4)

Species
C + ( L uC L ) = 0 t

(5)

Mass

+ ( Lu ) = 0 , t

(6)

where g is the liquid volume fraction and u is the volume averaged fluid velocity defined as: u = gu L Mixture Density

(7)

= S d + g L
0
Mixture Solute Density

1 g

(8)

(9) 0 where C is the volumetric specific heat, taken as volume fraction weighted averages. The boundary conditions at the domain are prescribed as: C L T u =0, and at z = 0 (10) k = hi ( T0 T z =0 ) =0 z z at z = Zb (11) T T p and C C0 The method used to determine the transient metal/coolant heat transfer coefficient, h, is based on the solution of the Inverse Heat Conduction Problem (IHCP) [17]. This method makes a complete mathematical description of the physics of the process and is supported by temperature measurements at known locations inside the heat conducting body. The temperature files containing the experimentally monitored temperatures are used in a finite difference heat flow model to determine h, as described in a previous article [18]. The process at each time step included the following: a suitable initial value of h is assumed and with this value, the temperature of each reference location in casting at the end of each time interval t is simulated by the numerical model. The correction in h at each interaction step is made by a value h, and new temperatures are estimated [Test(h+h)] or [Test(h-h)]. With these values, sensitivity coefficients () are calculated for each interaction, given by:
T ( h + h ) Test ( h ) = est h1

C = S C S d + g L C L ,

1 g

(12)

The procedure determines the value of h, which minimizes an objective function defined by:
n F ( h ) = ( Test Texp ) 2 , i =1

(13)

where Test and Texp are the estimated and the experimentally measured temperatures at various thermocouples locations and times, and n is the iteration stage. The applied method is a simulation assisted one and has been used in recent publications for determining h for a number of solidification situations [17]. 3. Experimental procedure The casting assemblies used in solidification experiments are shown in Figure 3. Considering the configurations designed, heat can be extracted only through the water-cooled system either at the top (Figure 3a) or at the bottom (Figure 3b), promoting vertical downward and upward directional solidification, respectively. A stainless steel split mold was used. The lateral inner mold surface was covered with a layer of insulating alumina to minimize radial heat losses. In both cases, it was employed a stainless steel connected part, with a wall thickness of 3 mm, to close the split mold. Previous works have carefully reported more details about the systems assemblage [7-8, 19].

(a)

(b)

Figure 3. Schematic representation of the experimental setups: a) downward and b) upward systems.

The alloys were melted in situ and the lateral electric heaters had their power controlled in order to permit a desired melt superheat to be achieved. To begin solidification, the electric heaters were disconnected and at the same time the water flow was initiated. A rotameter was used to control water flow. The device can promote water flow ranging between 0.84 and 0.9 cubic meters per hour. The experiments were performed with hypoeutectic Al-Si alloys (5; 7 and 9 wt% Si) with melt superheats of 5C above the liquidus temperature, and under a thermal contact condition at the metal/mold interface corresponding to the heat-extracting surface being polished. The heat-extracting surface was polished with grinding paper, and the surface roughness was determined with a digital system, where the arithmetic mean of the roughness amplitude profile (Ra in m) was adopted to characterize the surface microgeometry. In all cases, the chill surface roughness was kept constant, with a mean value of about 0.10 m. The employed thermophysical properties of these alloys are those reported in a previous article [19]. Casting temperatures were monitored during solidification via the output of a bank of fine type K thermocouples, sheathed in 1.6 mm O.D. stainless steel protection tubes, and accurately positioned in strategic positions with regard to the heat extracting surface. The thermocouples were calibrated at the melting temperature of aluminum exhibiting fluctuations of about 1.0 C. Thermocouples readings (at intervals of 0.5 s) were collected by a data acquisition system and stored on a computer.

4. Results and discussion Typical examples of experimental cooling curves concerning the thermocouples inserted into the casting, during downward transient directional solidification of an Al-7wt%Si alloy, are shown in Figure 4. The results of the experimental thermal analysis in castings were compared with simulations furnished by the finite volume heat flow program which was detailed in section 2, and an automatic search has selected the best theoretical-experimental fit from a range of transient heat transfer coefficients profiles, as described in a previous article [4]. Typical comparisons between calculated and experimentally measured curves can be seen in Figure 5. The curves refer to the upward transient solidification of an Al-9wt%Si alloy. Figure 6 shows the results which were obtained for the time dependence of metal/coolant heat transfer coefficient (hi) during the course of different experiments involving downward (Figure 6a) and upward (Figure 6b) directional solidification of hypoeutectic Al-Si alloys against uncoated water-cooled molds. It can be seen that the heat transfer coefficient, hi, rise with decreasing silicon content of the alloy for both experimental configurations examined. It can also be seen that lower, and essentially constants, heat transfer coefficients are associated with downward solidified castings. For example, in the case of the Al 5wt % Si alloy the resultant time dependent expression is hi = 2400 (t) -0.001 while for the vertical upward solidification of the same alloy larger values were obtained: hi = 4500 (t) -0.009. Thus, the heat transfer coefficient is clearly dependent on orientation of solidification with respect to gravity. A similar observation has been reported by Griffiths [9], who also found that upward solidification gave rise to larger values of hi than downward solidification of the same alloy. In upward solidification the effect of gravity causes the casting to rest on the mold surface, but during downward solidification, this action causes the solidified portion of the casting to retreat from the mold surface. Considering the former situation, the combined effects of volumetric shrinkage and casting weight results in the development of a considerable interfacial air gap, which is bigger than those typical of upward solidification. The reduction in the

contact pressure between casting and mold surfaces leads to a consequent reduction in heat transfer efficiency. The initial contact stage during directional solidification seems to be essential to the comprehension of the interfacial gap occurrence, since only few seconds may be sufficient in order to promote a solidified skin of the casting and its deformation. For example, the heat transfer coefficients for upward solidification are high in the initial stages of solidification, which result of the good surface conformity between the liquid core and the solidified shell. The mold expands while solidification progresses due to the absorption of heat and the solid metal shrinks during cooling. As a result, a gap develops because pressure becomes insufficient to guarantee a conforming contact between the surfaces. Once the air gaps forms, the heat transfer across the interface decreases rapidly and a relatively constant value of hi is attained, as shown in Figure 6b. Figure 6a shows constant values of hi along solidification. As the casting moves away from the chamber surface very rapidly due to casting weight during downward solidification, the sprouting of interfacial gap is faster than for upward solidification, which causes lower and constant hi values.
700

Al-7wt%Si
600

Temperature ( C)

500

400

300

200

100 0

Thermocouple 5 mm Thermocouple 13 mm Thermocouple 27 mm Thermocouple 41 mm Thermocouple 62 mm Thermocouple 90 mm 50 100 150 200 250

Time (s)
Figure 4. Experimental cooling curves for six thermocouples inside the casting directionally solidified vertically downwards. Curves are for thermocouples at 5, 13, 27, 41, 62 and 90 mm from the metal/cooling chamber interface. TP is the initial melt temperature.
700

Al-9wt%Si
600 500

Tp (Z) = -7550(Z) + 891.11(Z) + 874.65 [K]

Temperature ( C)

400 300 200 100 0 0 10 20 30 40 50 60 70 80 90 100 Thermocouple 4 mm Thermocouple 8 mm Thermocouple 12 mm Thermocouple 22 mm Numerical Simulation

hi = 3300 (t)

- 0,09

[W/m K]

Time (s)
Figure 5. Simulated and measured temperature responses at 4, 8, 12 and 22 mm from the metal/mold interface for an Al-9wt%Si alloy solidified upwards.

Metal/Coolant Heat Transfer Coeficient

4900

4900

hi = 2400 (t)
4200 3500 2800 2100 1400 700 0 20 40 60 80

-0.001 -0.001 -0.001

- Al-5wt%Si - Al-7wt%Si - Al-9wt%Si

hi = 4500 (t) hi = 3900 (t) hi = 3300 (t)

-0.09 -0.09 -0.09

- Al-5wt%Si - Al-7wt%Si - Al-9wt%Si


4200 3500 2800 2100 1400

hi = 2100 (t) hi = 1100 (t)

hi (W/m K)

100 120 140 160 180 200 0

20

40

60

80

700 100 120 140 160 180 200

Time (s)

Time (s)

(a)

(b)

Figure 6. Evolution of metal/coolant interface heat-transfer coefficient (hi) as a function of time (t) for Al-Si alloys during vertical (a) downward and (b) upward directional solidification. Figure 7 shows the temperature data collected in the metal during the course of upward solidification of an Al -3 wt%Si alloy casting in the vertical water-cooled apparatus with a melt superheat of 104C above the liquidus temperature. The experimental thermal responses corresponding to five different positions inside the casting were compared with the predictions furnished by the numerical solidification model. The best theoretical-experimental fit has provided the appropriate transient hi profile for two different approaches: (i) a mean melt temperature has been adopted (Fig. 3a), and (ii) a quadratic equation, based on experimental thermal readings, representing the melt temperature as a function of position in casting has been used (Fig. 3b). A comparison between hi profiles determined in each case is shown in Fig. 3c. It can be seen that a difference exists between the two curves, with the assumption of a constant melt temperature overestimating the metal/mold heat transfer coefficient. The two curves tend to separate from each other with increasing time, i.e., for thick castings the adoption of an initial melt temperature profile may be fundamental for providing accurate casting simulations.
800

800

700

700

600

600

Temperature ( C)

Temperature ( C)

500

500

400

400

300

Al-3wt%Si Tp = 758 C (mean) hi = 8500 . t


-0.23

200

[W/m K]

5mm 10mm

20mm 40mm

300

Al-3wt%Si Tp (Z) = -1161.5(Z) + 397.05(Z) + 1011.2 hi = 7800 . t


-0.20 2

100 0 50

70mm Numerical Simulation


100 150

200

[W/m K]

5mm 10mm

20mm 40mm

70mm Numerical simulation


100 150

100 0 50

Time (s)

Time (s)

(a)

(b)

6000 5500 5000

hi = 7800.t hi = 8500.t

-0.20 -0.23

[W/m K] [W/m K]
2

hi [W/m K]

4500 4000 3500 3000 2500 0 50 100 150 200 250

Time [s]

(c) Figure 7. (a) Simulated and measured temperature responses for Al-3wt%Si alloy at 5, 10, 20, 40 and 70 mm from metal/mold interface adopting a mean melt temperature; (b) Simulated and measured temperature responses for Al-3wt%Si alloy at the same positions adopting an initial melt temperature profile; (c) Evolution of metal/mold interface heat-transfer coefficient (hi) as a function of time for an Al-3wt%Si alloy casting. 5. Conclusions The following major conclusions are derived from the present study: 1. The transient metal/coolant heat transfer coefficients (hi) have been satisfactorily determined as a result of an approach based on measured temperatures along casting and numerical simulations furnished by a solidification model. 2. For both growth directions, the transient metal/coolant heat transfer coefficient, hi, can be expressed as a power function of time and rise with decreasing silicon content of the alloy. 3. Comparing both configurations analyzed and considering solidification of hypoeutectic Al-Si alloys, the determined hi profiles for upward transient directional solidification were significantly higher than those derived for the opposite condition. This is a clear indication that hi depends strongly on the direction of solidification, and that must be considered for an accurate simulation of freezing patterns in castings

Acknowledgements The authors acknowledge financial support provided by FAPESP (The Scientific Research Foundation of the State of So Paulo, Brazil) and CNPq (The Brazilian Research Council).

References 1. Ho K, Pehlke RD. Metall. Trans. B 1998; 16B: 585. 2. Krishnan M, Sarma DGR. Int. Comm. Heat Mass Transfer 1996; 23:203. 3. Santos CA, Siqueira CA, Garcia A, Quaresma JMV, Spim JA. Inverse Problems in Science and Engineering 2004; 12: 279. 4. Santos CA, Quaresma JMV, Garcia A. J. Alloys Comp. 2001; 319: 174. 5. Martorano MA, Capocchi JDT. Int. J. Heat Mass Transfer 2000; 43:2541. 6. Griffiths WD. Metall. Mater. Trans B 2000; 31B: 285. 7. Spinelli JE, Ferreira IL, Garcia A. J. Alloys and Comp.2004; 384: 217. 8. Spinelli JE, Rosa DM, Ferreira IL, Garcia A. Mater. Sci. Eng. A 2004; 383: 271. 9. Griffiths WD. Metall. Mater. Trans. B 1999; 30B: 473 10. Beck, JV. Int. J. Heat Mass Transfer 1970; 13: 703. 11. Loulou T, Artyukhin EA, Bardon, JP. Int. J. Heat Mass Transfer 1999; 42: 2119.

12. Piwonka TS, Woodbury KA, Wiest JM. Mater. Design 2000; 21: 365. 13. Voller VR. Can. Metall. Q. 1998; 37:169 14. Ni J, Beckermann C. Metall. Trans. A 1991; 22A: 349 15. Ferreira IL, Santos CA, Voller VR, Garcia A. Metall. Mater. Trans. B 2004; 35B: 285 16. Sundman B, Chen Q. STT Foundation, Stockholm, Sweeden, 1995 17. Santos CA, Quaresma JMV, Garcia A. J. Alloys Comp. 2001; 319: 174 18. Krishnan M, Sharma DGR. Int. Comm. Heat Mass Transfer 1996; 23: 203 19. Peres MD, Siqueira CA, Garcia A. J. Alloys Compd. 2004; 381: 168

Das könnte Ihnen auch gefallen