Sie sind auf Seite 1von 16

HEAT TRANSFER OF LAMINAR FLOW COOLING DURING STRIP ACCELERATION ON HOT STRIP MILL RUNOUT TABLES

Remn-Min Guo Sr. Staff Engineer Dept. of Automation Technology ARMCO Research & Technology ABSTRACT Runout table temperature control is one of the most critical processes to ensure mechanical properties of steel strip. Accelerated cooling has been implemented in hot strip mills to reduce the heat loss on the delay table, to achieve the required coiling temperature and to obtain the desirable grain structure. Appropriate cooling application with better understanding of strip cooling characteristics guarantees the precise temperature and cooling rate control. This article describes a model which determines heat transfer coefficients using operating data and an inverse technique. INTRODUCTION Immediately after the strip rolls out from the last stand of the finisher, it enters the runout table cooling zone. Depending on the finishing temperature, mill speed, and the required coiling temperature, the material is cooled by arrays of top and bottom cooling headers in a distance of about 70% of the runout table length, and continuously cooled by air convection and radiation until it is wound in the downcoiler. The symmetric cooling effect of the last dry cooling section allows the material to re-distribute temperature by internal conduction. It is particularly important for mill operation with inevitable non-symmetric cooling in the water cooling zone[1]. This cooling process for the strip to cool from approximately 870oC to below 730oC. It includes internal and external conduction, stagnation forced convection, forced boiling convection, air convection and radiation, and heat generation from material phase transformation (see Figure 1). Although table roller conduction was recently found to be also a cooling element to cause a large local temperature drop [2], external conduction as well as air convection are usually treated together to obtain an empirical heat transfer coefficient equation for the mill application [3,4]. Radiation heat transfer follows Stefan-Boltzmann rule with temperature dependent emissivity [5]. Heat generation due to phase transformation depends on the ratio of transformed austenite. Transformation detectors have been applied to the runout table temperature control [6]. Nevertheless using temperature dependent thermal properties reduces the problem complexity and is widely adopted and accepted in the literature.[1-5, 721]. Forced convection by the cooling medium which counts for more than 90% of the entire heat transfer is still not fully understood because heat transfer mechanism involves a complex mixing phenomenon of water impingement and boiling on a moving surface. For a production mill, analyzing the operating data

can hardly obtain a concrete conclusion because of inherent data error from the mill harsh environment and the confounding effects between rolling parameters. Investigations using a small scale equipment can provide more information, however these results do not reflect the true mill operation and call for modifications before application. In order to apply these academic knowledge to mill practice, it is useful to review the previous research works of stationary, steady state moving, and mill application cases in the following. Stationary Cases For a steady state condition of water impingement on a stationary plate, heat transfer adjacent to a jet impinging was roughly classified into five regimes by Zumbrunnen [22]. As shown in Figure 1, the single phase forced convection zone is located right beneath the jet. Followed by the wet zones of nucleate/transition boiling, forced convection film boiling, and agglomerated pools, heat transfer gradually reduces to the dry zone by air convection and radiation. The impingement area was estimated about 2 to 4 times of the nozzle diameter or the curtain width[12]. At the vicinity of the impact zone, an observable darken line [23] shows that this zone may have the highest heat transfer. Otomo's [24] experiment showed that heat transfer coefficient at this stagnation area held constant (23.26 kw/m2-oK) regardless of the impinging speed and the surface temperature if the surface temperature exceeded 400oC. Experiment conducted by Ishigai demonstrated that the cooling surface temperature near the stagnation point was cooled from an initial temperature of 1000oC to 180oC in 2.5 sec, an average cooling rate of 328oC/sec [25]. From his boiling curve the transition and nucleate boiling temperature was found around 675oC and the maximum heat transfer occurs at 300oC. Using various impingement angle from a spray nozzle, Ellerbrock [13] concluded that heat transfer sharply increases once the surface temperature passes the Leidenfrost point, which is near 670oC although varies from test to test. It was also reported that heat flux of the nucleate boiling is not a function of the impinging speed and the plate surface temperature [25]. Summarizing from these results, the stagnation zone does have highest heat transfer effect which depends on the surface temperature and the boiling condition. Heat transfer is relatively small at the onset of water impingement on the plate because film boiling prevails and no audible boiling noise (no active boiling). After a certain time delay which depends on the experiment, heat transfer rapidly raises as the surface temperature reduces to change film boiling to transition and nucleate boiling. Away from the impingement zone, an "effective cooling zone" can be visualized about 20 to 48 times of the nozzle diameter depending on the jet Reynold's number [14]. Since the wet zone has higher heat transfer ude to cooling medium, it is named the effective cooling zone henceforward in this article. The local and average heat transfer of the effective cooling zone was reported by Zumbrunnen as a function of Reynolds number and the distance from the stagnation point [26]. For a constant jet speed, heat transfer coefficients in the effective cooling zone are nearly linear to the distance even the regression equations which he used have higher order terms. Moving Cases Many studies were also performed for the steady state condition with a moving plate. The impact zone length remains the same as the stationary case, but the wet cooling zone length stretches downstream to 19 times of the nozzle diameter and only 3 times upstream[14], which makes a non-symmetric effective cooling zone. The cooling mechanism appears similar to the stationary case. The highest heat transfer occurs at the regime of nucleate boiling and the surface temperature and the plate speed has a negligible effects at that state. The heat transfer coefficient is about twice of film boiling regime [14]. However, the nucleate boiling occurs earlier than the plate reaches the stagnation point and causes the peak heat transfer to shift to the direction opposite to the plate motion. Furthermore, the heat transfer coefficient seems symmetric to the maximum plane and shows a similar trend as observed in the stationary case [22].

Obviously, the maximum heat transfer location appears to be at the vicinity of the onset point of nucleate boiling instead of the stagnation point. However, both the Lindenfrost temperature and its corresponding location are difficult to formulate theoretically. In fact, the Lindenfrost point and the heat transfer coefficient are highly correlated each other. Namely, they depend on cooling capacity of water flow [13] as well as the initial temperature [24]. In his early study, Auman first pointed out that the surface temperature of 538 - 649oC represented the nucleate and transition boiling temperature using various initial temperature [15]. Zumbrunnen selected an initial temperature of 644oC and 108oC, and found that nucleate boiling cannot occur in the low initial temperature [22]. Using slightly higher initial temperature of 240oC, Han et. al. found the nucleate boiling regime is in the range of 120oC - 170oC.[14] Since water saturated temperature is 100oC, the low initial surface temperature in the impingement zone might be too cool for boiling. A higher initial temperature will cause nucleate boiling with a time delay which depends on the water cooling effect. Operating Cases Controlling coiling temperature in the production mill has to take additional operating conditions into consideration. Since strip gauge, finishing temperature, mill speed, and cooling headers vary during rolling, the mill hardly reaches a steady state condition. Low pressure, high volume cooling headers are activated or deactivated every second with a large response time. The laminar jet speed and its flow rate are in the neighborhood of 1 m/sec and 22 liter/min (lpm) respectively. Since the mill speed is in a range of 2.5 to 15.3 m/sec, the plate to jet speed ratio of 2.5 to 15.3 is much larger than the experimental cases of less than 0.5. Header cooling effect is substantially reduced due to cooling water interaction if two adjacent headers are open simultaneously. Because of high volume, water stays on the strip and causes washback and washdown effect. Water washback effect can be realized by the reported problems of the X-ray gauge meter error due to water washback from the first cooling header, although the distance from the first header to the F7 stand is as far as 15 meter (see Figure 2). Similarly, water washdown effect is observable that the residual water is usually blown away by high pressure water(or air) wipers mounted at both sides of the runout table. Unlike the bottom header, water drops when it lost its momentum to the opposing gravity force, the large amount of cooling water from the top header can not escape from the strip surface in a very short time. Except a portion of water evaporates, most of water washes downstream or upstream and gets off the strip from the strip edge. Particularly for a wider strip, the water in the central portion of the strip has more difficulty to escape from the strip edge. Consequently, it is very difficult to apply the experimental result to mill operation directly. Alternative approaches adopted in the steel industry is to utilize statistical methods to obtain an empirical equations to calculate the required headers for cooling temperature control [16,17]. Thanks to the feed forward and adaptive control loops, many satisfactory results were reported [18,19,20]. Simple and low maintenance are the advantages of this approach. The disadvantages are difficulty of fine tuning the controller and the frequencies of re-tuning the controller when changing rolling schedules. The objective of this study is to modify heat transfer mechanism experienced from the lab to best suit for the mill application.

DATA COLLECTION AND DATA UNCERTAINTY Data for a total of 75 coils were collected in 1990 from an 86" hot strip mill of Middletown Works, ARMCO Steel Company, L. P. (ASCLP). This includes 39 coils without bottom header cooling and 36 coils with both top and bottom header cooling. A brief layout of the runout table is illustrated in Figure 2. The X-ray gaugemeter and the pyrometer for finishing temperature are mounted right after the F7 mill stand. The specification of the mill is listed below: rolled material: finishing temperature: coiling temperature: mill speed: strip gauge: strip width: coil length: low carbon, Ti-IF, high carbon and HSLA, 843o - 927oC, 594o - 743oC, 2.53 - 15.23 m/s, 1.80 - 9.53 mm, 635 - 2032 mm, 79 - 1417 m.

Data was collected in a real time basis of every one second snapshot. Each coil data includes top header on/off condition and instantaneous water pressure of bottom headers, strip gauge, F7 work roll speed, finishing temperature and coiling temperature. Figure 3 shows typical rolling data during mill acceleration for coils #28 and #29. The sources of data uncertainty include the sensor error, the response of the actuator, and the disturbance of mill operating environment. The finishing and coiling temperature were measured by an infrared type pyrometer with 1% of accuracy. This leads to about 9o and 7oC error of the finishing and coiling temperature respectively. Other uncertainty sources promote temperature error, such as water washback effect, residual water, vapor fume, strip shape and bow, strip flutter, and strip walking. An estimated temperature error could be as high as double of the manufacturing specification depending on mill operation conditions. Finishing temperatures were measured by a top mounted pyrometer while coiling temperatures of the bottom surface was measured by a bottom mounted pyrometer to reduce the error due to residual water. The strip gauge was measured using a Weston's X-ray gaugemeter. During rolling, the sensor aims to the strip center for central gauge measurement and control. The standard accuracy is 0.1% and the typical noise is 0.1% at 30 ms response time. The mill speed was measured by a tachometer of the #7 finishing stand work roll for most part of the coil. When strip tails out from the #7 stand, the speed data was not recorded and the theoretical computed speed was reported instead. This portion of data was excluded for model identification and verification because of possible large error of mill speed. Since the exit strip speed is higher than the roll speed, element tracking by the roll speed causes error due to forward slip. The maximum pumpage for the 40 top cooling headers (1.37 m apart) are 5548 l/m for four pumps. Top headers of laminar flow tube type with constant flow rate are controlled in pairs by two 2-byte integers to activate diverting/cooling toggle valves. It was measured about 2 seconds response time for water to contact the strip since the signal had issued by the controller. The 18 bottom headers are single wide-angle flat jet nozzles spraying from 740 mm below the passline. Experiments show that flow rate was lower at the center and reaches the maximum at about 840 mm from the center regardless of water pressure[13]. Top and bottom headers were designed to activate in any combination. With the non-symmetric header arrangement and the current control algorithm, the runout table has been operated in a non-symmetric cooling fashion [1]. Since the coiling temperature differences from the center to the edges were recorded as high as 55oC [13] due to non-uniform cooling of bottom headers, this study will concentrate on the top

header cooling calculation using collected 39 coil data without bottom header cooling. A follow-up study will include the bottom header cooling using heat transfer coefficient equations obtained from this study. HEAT TRANSFER EQUATIONS AND MATHEMATICAL MODELS Mathematical models are developed using the Lagrangian coordinate system to track the strip element from the finisher to the downcoiler. The heat transfer equation can be expressed as: (1) where = material density c = specific heat, a function of T T = strip temperature t = time k = conductivity, a function of T L = latent heat of transformation Z = ratio of transformed austenite. Heat transfer in the rolling direction (x) and in the width direction (z) are substantially smaller than in the thickness direction (y) due to the high rolling speed [21] and uniform cooling capability of the header respectively. Accordingly, Eq (1) is simplified to a one dimensional unsteady state equation [1] and its corresponding initial and boundary conditions for the top and bottom strip surfaces are t = 0 T = Tf T t >0 k = h(T Ta ) y T Tw T 4 Ta4 h = ha + hw + T Ta T Ta Tf = finishing temperature h = equivalent heat transfer coefficient ha = heat transfer coefficient of air convection hw = heat transfer coefficient of water convection Ta = ambient temperature Tw = water temperature y = thickness direction = Stefan - Boltzman Constant = steel emission coefficient. c T dZ = kT + L dt t

(2)

where

Transformation heat L is an important heat source in the runout table cooling zone where the transformation occurs. Since Z is a function of strip temperature and material property, inclusion of transformation needs an additional iteration loop to check dZ/dt which will substantially increase computing time. Even with the transformation detector, the transformation ratio is assumed a complementary exponentially decay function in Yashiro's model instead of using implicit calculation [6]. Yet since most production hot strip mills do not have transformation detector, without measured data, dZ/dt will become an adjustable variable for the model if L is taken into consideration. This will improve model accuracy because of additional adjustable variables, take longer time for model tuning, and also reduce generality of the model. Most studies neglect the latent heat of transformation L and employ the temperature dependent thermal properties[1-5,7-21]. The same approach is adopted in this article and selected thermal properties were developed by Pehlke, et. al. [27]. Further simplification of Eq. (1) using the lump system, an assumed temperature profile, and the implicit finite difference method can be found in Guo's works and will not reiterate here [1,28]. The results show that the finite difference method has the least standard error of forward calculation and the lump system has the greatest. However, the error difference is small - in particular for thin gauge strip - but the computing time is significantly different. This article follows the implicit finite difference method. The equation is solved by applying moving boundary conditions to fixed strip elements. The heat transfer coefficients of air convection are basically derived from Thomas's formula [4] with a modification factor Ka: ha = 0 . 010093 K a v 0.926 , where v is strip speed in m/sec. Data from an additional coil without water cooling is used to determine the modification factor Ka. The tuning procedure is so simple (one variable and one coil data set) that no further description will be reiterated here. The heat transfer coefficient equations of water cooling are assumed as a function of coolant flow rate and strip speed, surface temperature and thickness. Data with top header cooling is used to estimate the power exponents of the presumed heat transfer coefficient equation. The effective cooling zone of each cooling header (later change to the effective cooling time for computation convenience) is determined so as to obtain the least square error of the heat transfer coefficient equation. Model verification is performed by the forward calculation using identified heat transfer coefficient equations. HEAT TRANSFER COEFFICIENT FOR TOP HEADERS Each top header consists of 87 nozzles which are arranged in two rows spacing 82.6 mm in the rolling direction. The front row includes 44 nozzles with 50.8 mm nozzle spacing in the transverse direction. The downstream row has the same arrangement but side shift 25.4 mm transversely to the front row. The 235 mm nozzle length facilitates to produce a laminar flow through a 20.64 mm ID bronze tube. The laminar flow header is designed to serve high volume and low jet speed water cooling. The nozzle capacity is approximately 2.5 l/m at 3447 pa static pressure. Due to pumpage capacity, the flow rate is limited to 1.6 l/m per nozzle, which yields a maximum Reynolds number Rej of 23,500. The impingement zone of a single nozzle is about 41.3 - 82.6 mm. Since two impingement zones overlaps, the heat transfer coefficient distribution is considered using the header as a unit. Practically speaking, it takes only 8 ms for a steel element to travel between two nozzles with an average mill speed of 10.2 m/s. Consideration of each nozzle leads to select a very small time interval, and small thickness division to ensure numerical stability. Furthermore the headers are controlled in pairs and cooling interaction between these two headers is indisputable. It seems reasonable to consider two headers together as one unit to further reduce the computing time. If this is the case, the question is what type of heat transfer coefficient distribution should be applied. The maximum heat transfer occurs not at the stagnation points for the moving strip [22]. Adding the header cooling interaction, heat transfer coefficient distribution is even more laborious to

determine. It may have only one peak point, but the possibility of twin peaks cannot be ruled out. Unfortunately, there are lack of basic studies concerning heat transfer behavior of an array of water jets impinging on a moving plate. Numerical solutions on the cooling effect of two submerged air jets revealed that Nusselt numbers for the stationary plate increase sharply with Reynolds number at the geometric midplane where the adjacent jet flows merge [29]. The maximum Nusssel number was not located at the stagnation point but off the nozzle opening toward downstream for a moving plate. Since the cooling behavior is not the same as the water cooling cases using a single nozzle, the distribution of heat transfer coefficient cannot be inferred accordingly. Two stage heat transfer coefficients were applied by Colas, et. al. [12]. Combining with the assumed oxide layer thickness and impingement length, their simulation showed that in the impingement zone an approximate 270oC temperature drop of the oxide film caused the appearance of a black band. The metal surface temperature never reaches below 650oC and its local temperature drop is only 10oC. However, their assumption of a delta function type heat transfer coefficient distribution contradicts with Zubrunnen's findings of a polynomial distribution [22]. For a steel element traveling from the finisher to the downcoiler, the total heat transfer from the rolling data is the same regardless of the assumed distribution of heat transfer coefficient. Reduction of the length of the effective cooling zone (see Figure 1) certainly needs to increase the maximum heat transfer coefficient to compensate for the heat loss, and vice versa. Since most of the runout table is exclusively confined in film boiling mode [15], the heat transfer ratio within the effective cooling zone cannot be as large as 110 times estimated by Colas. A reasonable ratio should not exceed 4 [14,15,25]. Combing the above findings, a simple but effective assumption for the average heat transfer coefficient distribution is made. The highest average heat transfer occurs at the center plane between these two headers and linearly decreases in both upstream and downstream directions until the dry cooling zone. This assumption simply extends Colas' delta function to a triangular distribution which is closer to Zumbrunnen's curve[22]. The maximum heat transfer coefficient (height of the triangle) and the effective cooling length (width of the triangle) are to be determined by the inverse method. Note that the effective cooling length is a function of strip speed [11]. The effective cooling time as shown in Figure 4 was adopted to take advantages of easy tracking and automatic speed dependent cooling length. A drawback of this assumption is that the location of the maximum heat transfer is not necessary to be at the center plane and the effective cooling zone may not be equal on both sides. Further inclusion of these two additional variables - the location and the upstream cooling length - in the study will certainly improve the results.

INVERSE METHOD The inverse method was applied to backward calculate the heat transfer coefficient equation using the optimal effective cooling time. Some of the computing procedure have been described in detail in reference 1. Basically, this scheme is to track each steel element from the finisher to the downcoiler. In the backward computing procedure, the on/off time history for each header can be generated using two control bytes. Consider a steel element whose temperature is to be determined. Its arrival time at each pair of headers is calculated using the mill speed chart which is available in the data. The heat transfer coefficient for this element can be estimated considering the header response time, the element arrival time, and the header on/off conditions. The relative maximum heat transfer coefficient is set to be 1. The absolute heat transfer coefficient can be determined by the shooting method using the known finishing and

coiling temperature and the moving boundary conditions which is discussed in detail in the following section. Heat transfer coefficient equation The absolute maximum heat transfer coefficient of each pair of headers is statistically determined by the following equation: (3) where K = curve fit constant v = strip speed v0 = reference strip speed t = strip thickness t0 = reference strip thickness Ts = strip surface temperature Ts0= reference strip surface temperature q = header flow rate q0 = reference header flow rate a,b= exponent of strip speed and thickness c,d= exponent of surface temperature and flow rate. The exponent of for both pipe (laminar tube) and slit (water curtain) nozzle headers were found about 1 [24], namely the average heat transfer coefficient is linearly proportional to the water flow rate in a range of 0.4 - 0.9 m3/min-m2. The top header flow rate is constant after response time and is proportional to time in the transient state. Consequently, the exponent d=1 is assumed and the steady state flow rate is used as the reference flow rate q0. A typical time history curve of water cooling relative heat transfer coefficient is also shown in Figure 4. The left most header shows that the heat transfer coefficient reduces in proportion because the flow valve is still not fully open when the strip element arrives at the header. Some following headers are close and provide zero heat transfer coefficients. The next two pairs of header have cooling interaction because their effective cooling length cover each other. No superposition effect in the interaction zone is assumed. Note that the maximum heat transfer coefficient of each header is not the same, instead, depends on the instantaneous surface temperature and strip speed. This time history curve of the average heat transfer coefficient forms the moving boundary condition. Every node points in Fig. 4 are considered and the computing time intervals between nodes are determined after some preliminary test runs. Determination of computing time interval and thickness division There are two time intervals - one is for element tracking and the other is to calculate temperature for each element using finite difference method. The tracking time interval of 1 sec was selected in the identification stage to match with data collection frequency of 1 Hz. Since the Lagrangian moving frame is employed and heat transfer along the rolling direction is neglected, the tracking time interval doesn't affect v hw = K v 0
a

t t 0

Ts T s0

q q 0

the result but the computing time interval does. Coiling temperature profiles for various computing time intervals and thickness divisions are plotted in Figure 5. For the computing time interval of 1 sec, the total calculation times are 0.49, 0.77, 0.99, and 1.32 seconds for thickness divisions of 11, 51, 101, and 149 respectively. As to 0.1 sec computing time interval, the total calculation time increases to 2.03, 3.13, 4.51, 5.88 seconds respectively, which is about 4 times of the previous case. A larger computing time interval leads to smaller temperature and less total calculation time. Thick divisions of 11 and 51 have very close results regardless of computing time intervals. The temperature difference is less than 1oC within the same computing time interval group, and is about 1.5oC for all curves, which are all smaller than measurement errors of 7oC. In the identification stage, using the small computing time interval and large thickness divisions costs much more total calculation time. Selection of 11 thickness division and 1 sec computing time interval satisfies most strip elements and produces a stable numerical solution. Two stage model tuning Model tuning includes two major steps - backward model identification and forward model verification. The backward identification is to estimate a,b, and c of Eq (3) using curve fit technique and collected 39 coil data. The forward model verification is to examine the goodness of model prediction on coiling temperature using the same rolling data. For a strip element to travel on the runout table, strip gauge is invariant. Strip speed changes to some extent depending on mill acceleration. Surface temperature varies rapidly and unevenly, especially in the cooling zone. Using the average surface temperature of each coil to fit Eq (3) leads to a very large forward verification error even the curve fit error of the heat transfer coefficient equation Eq (3) is quite small. The same but less severe situation occurs when using the average strip speed of each coil for Eq (3). One effective way to offset this problem is to fit Eq (3) in two stages. Figure 6 describes the overall flow chart including the estimation loop and the optimization loop. The estimation loop is employed to obtain an estimation of exponents of a, b, and c. Since strip gauge is constant, the exponent b calculated from the curve fit technique does not disturb the forward calculation. In the proximity of variables a and c, the optimization loop selects a combination of a and c first. The backward calculation is processed and the results are used to fit only b since a and c are predetermined. The forward calculation is then followed to carry out the forward error. The backward and forward errors are evaluated as a reference for next selection of a and c. The above procedure is performed for various effective cooling time. Identification and verification results Figure 7 shows the forward and backward errors for various effective cooling times. The minimum backward error of 0.085 kw/m2-oK (7.3%) occurs at the effective cooling time of 1 sec, and the least square error of 6.6oC at 0.9 sec. Figure 8 compares the heat transfer coefficient hw calculated from the backward identification stage and Eq (3) of the curve fit equation using 1 second effective cooling time. The benefits of forward coiling temperature prediction can be observed from Figure 9 which shows the measured and predicted coiling temperature using 0.9 second effective cooling time. Both two figures embrace 1050 data points and strip gauge ranges from 2.11 to 9.42 mm, rolling speed from 3.8 to 15.2 m/s, finishing temperature from 854o to 911oC, and coiling temperature from 649o to 727oC. Note that the standard error due to data uncertainty of two pyrometer is 3.8oC (3=11.4 oC). The forward error as calculated by the model is considerably satisfactory considering mill harsh environment and other data uncertainty as described before. Within the mill speed range, the 1 sec effective cooling time translates into 2.4 - 24 meter effective cooling length or 110 - 740 times of nozzle diameter. This result agrees with Filipovic's observation of 450 jet widths [2]. For the effective cooling time greater than 1 second, the

backward error increases slowly while the forward error increases rapidly. After careful evaluation, it was decided to select the effective cooling time of 1 second, 0.8 for the exponent a, and 1.4 for c. Since the strip gauge does not change during cooling, the term k(t/to)b of Eq (3) is modified to a polynomial function of t/to. The bottom chart of Figure 3 shows measured (solid line) and predicted (dashed line) coiling temperature for coils #28 and #29. SIMULATION RESULT Steady State The steady state rolling seldom occurs in a production hot mill. For thin gauge strip such as 2.54 mm, the mill speed reaches 10.2m/s when the head end leaves the finisher and gets into the runout cooling section. Figure 10 shows the isothermal lines for 2.54 mm strip with 871oC finishing temperature and 10.2 m/s constant speed. The top line represents the strip top surface and the bottom line does the strip bottom surface. The inversed triangle sign indicates the open header pair #10 location. The right most line at 147.15 m represents the pyrometer location at the downcoiler side (see Figure 2), i.e. the end of the runout table. The same definition is also applicable to Figures 11 - 15. The largest temperature gradient along the rolling direction is observed before the header pair. This agrees with Han's finding[14]. The isothermal line of 782oC breaks into two parts, one forms a closed loop beneath the top strip surface and the other runs through the strip thickness in the latter location. It is because the local surface cooling cannot penetrate through strip thickness in a short time. A similar phenomenon is found in the higher strip speed of 15.2 m/s as shown in Figure 11. The bottom portion of isothermal lines near the cooling zone lean toward downstream to satisfy the bottom boundary condition. For a slower speed of 5.08 m/s (not shown), the loop disappears and every isothermal line crosses the strip thickness with very small distortion because the small speed provides enough time to re-distribute temperature by virtue of internal conduction. For the case of 7.62 mm gauge, Figure 12 illustrates many semi-elliptic isothermal loops concentrating right after the cooling zone. The cooling effect has more difficulty to diffuse through a thicker material even the mill speed is slower. The center portion of the isothermal line moves toward downstream, which implies the warmer center of the strip cross section. As shown in Figure 13 the higher speed of 15.2 m/s generates more loops at the cooling zone. The loops span wider due to the larger effective cooling length. As the strip thickness and rolling speed increases, the local top surface temperature drop increases, but the overall coiling temperature drop decreases. Accelerating state Most rolling mills start to accelerate as soon as the head end enters the first finishing stand to prevent further heat loss on the delay table during rolling. The acceleration ranges from 0.025 to 0.072 m/sec2 depending on the material grade, strip width, total reduction, work roll size, the available motor horse power, and so forth. A very short period of time of constant speed may exist when the mill reaches the preset top speed. However, this short period cannot be qualified as a steady state rolling due to strip gauge and finishing temperature changes during this period. Consider one strip element just entering the runout cooling zone. The travel time to arrive at the downcoiler is usually greater than the short period of constant speed. Study on temperature distribution of the accelerating state should track on many strip elements with small tracking time. Computation can be suspended at any selected time. The location and temperature distribution of all elements on the runout table can then be recorded using the interpolation method. This "time_frozen" technique provides a method to investigate temperature distribution during mill acceleration. Obviously, this technique requires smaller computing time interval for better accuracy.

10

In this section, 0.1 second computing time interval is employed to track the element and also to calculate temperature distribution. The mill was assumed to accelerate when the head end passes the last finishing stand. It is reasonable to utilize an acceleration of 0.051 m/s2 for 2.54 mm gauge to accelerate from 10.2 to 15.2 m/s, and 0.025 m/s2 for 7.62 mm gauge from 5.08 to 6.72 m/s. Previous discussion shows that mill speed has a large effect on temperature distribution. Since mill acceleration takes most of the rolling time, temperature distribution for any specified time varies according to strip speed. Two particular time points as listed in Table I were selected for investigation. Note that the assumed mill speeds and rolling times as shown in Table I may yield a larger-than-usual coil PIW (Per unit Inch Weight), in particular, for the thick gauge case. Figure 14 and 15 show the results of the constant acceleration state. Isothermal lines of the high and low speeds are placed together for comparison. Temperature distribution is similar to the steady state cases. Particularly, these isothermal lines of the accelerating state are close to their steady state counterpart of the matching average mill speed (not shown). Two isothermal lines for high and low speeds are closer in thin gauge case. The curvature is getting larger at the end of the runout table as the mill accelerates. This makes the temperature non-uniform across strip thickness, in particular, thick material. As time elapses, isothermal lines of the high speed lead that of the low speed because the increasing speed deprives of the cooling time and raises the coiling temperature. However, there is an opposite trend in a small area near the strip surface before the stagnation point. This is because the local heat transfer increases as the mill speed increases. And when the strip element approaches the stagnation point, the cooling effect due to the increasing local heat transfer coefficient makes that local area cooler in the high speed condition. Cooling Comparison between Steady and Accelerating State The travel time of the strip element needs only 10 sec starting from the finisher at 14.73 m/s arriving at the downcoiler at 15.24 m/s. Figure 16 shows temperature distribution for two strip elements of 7.62 mm gauge as a function of its travel distance from the finisher using cooling header pairs of #1, #5, #10, #15, and #20. Element A starts at 5.08 m/s and accelerates to 5.77 m/s and element B runs a constant speed of 5.43 m/s (average speed of element A). In the cooling zone, temperature distribution has nearly no differences in strip center and bottom surface. Temperature gaps at the top surface do exist due to speed difference. Element A travels slower than B in the first half of the runout table and faster in the second half. Its temperature is slightly warmer at the first 70 m, and becomes slightly cooler after that. Temperature difference even becomes smaller for the thinner gauge cases. It concludes that for each strip element, one can use an average steady state speed to estimate temperature drop for an accelerating state. This facilitates the further development of the control algorithm. The cooling rate plays an important role in material grain structure. The average cooling rate ranges from 5 to 25 oC/sec depending on the mill speed and the finishing and coiling temperatures. The local cooling rate is usually measured indirectly. The inverse methods are applied in the lab to back calculate the cooling surface temperature from the measured temperature. The temperature-time curve can be used to estimate the local cooling rate. The maximum cooling rate was found about 1680 oC/sec by using Ishigai's data [25] and 120oC/sec by Otomo [24]. The strip surface is expected not to drop tremendously because the fast strip speed reduces the cooling time. And the stationary cases of their experiment have a higher heat transfer than the moving cases of the mill. An estimated strip travel time of a production mill is about 0.0061 second using an average mill speed of 10.2 m/s and impact zone of 62 mm (3 times of nozzle diameter). The maximum local temperature drop should be in a range of 0.73 to 10oC. The instantaneous

11

cooling rates shown in Figure 17 are calculated from Figure 16 using the Stirling formula. The maximum cooling rate of the top surface is about 300 oC/sec at the first cooling header pair. Large local cooling rate and temperature drop may quench the strip surface to form oxide film which was considered in Colas's model [12]. Heat Transfer and Cooling Efficiency The overall heat transfer (kcal/m2) is calculated by integrating the local heat flux over the element travel time. Figure 18 shows that the overall heat transfer decreases as the mill speed increases. Even the local heat transfer coefficient increases for the higher speed, the shorter exposure time overrides that effect and reduces the overall heat transfer. The higher finishing temperature has a larger heat transfer coefficient and in turn a larger overall heat transfer. The cooling efficiency can be evaluated by heat removal per unit mass of cooling water. The mass flow rate for the case of all cooling header open is 1325 kg/sec. Figure 19 compares the cooling efficiency for various mill speed, finishing temperature and mill acceleration. It is obvious that increases of mill speed, finishing temperature, and mill acceleration lead to better cooling efficiency. The cooling efficiency is smaller than Chen's findings [29] and closer to Otomo's [24]. CONCLUSIONS Many studies have been proceeded in the impingement cooling area using a small scale laboratory equipment. Important factors in the production mill were frequently neglected however, such as mill speed changes, gauge variation, cooling header on/off timing and duration, and cooling header interaction. Other studies on heat transfer of strip cooling for the practical application relied on statistical methods more than analytic. Consequently, the equations derived from the laboratory cannot be applied directly to the mills and the statistical equations are suitable for mill operation but lack of physical meaning. This article describes an inverse method to obtain the heat transfer coefficient equations using the operating data and valuable knowledge from the experimental works. A triangular shape distribution of heat transfer coefficient was assumed. Consideration of tracking time, header on/off time, and header response time expedites a better prediction of the model. The effective cooling time of 1 sec provides the best approximation. While the effective cooling time pronounces its importance, the impingement length seems too small to affect the overall runout table cooling. The backward error is 7.3% based on a simple exponent equation for a total of 1050 data points. Its corresponding forward error is only 6.78oC with temperature data uncertainty of 3.8oC. The average heat transfer coefficient has positive association with strip gauge, strip speed, and finishing temperature. The thicker gauge leads the larger heat transfer coefficient, which may result from faster surface temperature recovery. Higher mill speeds provide greater heat transfer coefficients, deeper cooling penetration, better cooling efficiency, but smaller total heat transfer. The warmer finishing temperature gains better heat transfer due to larger heat transfer coefficient and temperature difference. For the accelerating state within normal mill acceleration, the average speed of each strip element can be used to estimate total temperature drop with an acceptable error. This finding provides a great convenience for the runout table cooling control. ACKNOWLEDGMENTS The author like to thank the technical staff of the Hot Strip Mill of ASCLP for collecting data and mill information, the rolling group of Automation Technology of ARMCO Research & Technology for helpful discussion, and Mr. Ellerbrock for his valuable and in-depth suggestion.

12

Table 1: Top speed time and low speed time Thin Start Speed End Speed Low Speed Start Time End Time Average Speed Start Speed End Speed High Speed Start Time End Time Average Speed 10.16 10.87 0 14 10.52 14.73 15.24 90 100 14.99 Thick 5.08 5.77 0 27 5.42 7.11 7.62 80 100 7.36

13

FIGURES: Figure 1: Heat transfer regimes adjacent to a jet impinging on hot strip Figure 2: Runout table layout of the 86" Hot Strip Mill, Middletown Works, ARMCO Steel Company, L. P. (ASCLP) Figure 3: Collected rolling data of finishing temperature, F7 mill speed, F7 exit gauge, and coiling temperature for coils #28 and #29. Solid line: Dashed line: measured coiling temperature, predict coiling temperature

Figure 4: A triangular shape of the average relative heat transfer coefficient as a function of time Figure 5: Coiling temperature profile of 2.54 mm strip for various calculations; labels show the thickness division and computing time interval Figure 6: Flow chart of inverse calculation and estimation & optimization loops Figure 7: Backward and forward errors for various effective cooling time Figure 8: Comparison of heat transfer coefficient using 1 sec effective cooling time Figure 9: Comparison of measured and predict coiling temperatures using 0.9 sec. effective cooling time Figure 10: Isothermal lines for 2.54 mm gauge, 871oC finishing temperature, 10.16 m/s speed, and the 10th header pair ON Figure 11: Isothermal lines for 2.54 mm gauge, 871oC finishing temperature, 15.24 m/s speed, and the 10th header pair ON Figure 12: Isothermal lines for 7.62 mm gauge, 871oC finishing temperature, 5.08 m/s speed, and the 10th header pair ON Figure 13: Isothermal lines for 7.62 mm gauge, 871oC finishing temperature, 7.62 m/s speed, and the 10th header pair ON Figure 14: Isothermal lines for 2.54 mm gauge, 871oC finishing temperature, and the 10th header pair ON; solid lines: accelerates from 14.73 to 15.24 m/s; dotted lines: accelerates from 10.16 to 10.87 m/s Figure 15: Isothermal lines for 7.62 mm gauge, 871oC finishing temperature, and the 10th header pair ON; solid lines: accelerates from 7.11 to 7.62 m/s; thin lines: accelerates from 5.08 to 5.77 m/s Figure 16: Temperature distribution on the runout table, solid lines: accelerates from 5.08 to 5.77 m/s (Element A); dotted lines: constant speed of 5.42 m/s (Element B) Figure 17: Temperature change rate for 7.62 mm gauge, 871oC finishing temperature, solid lines: accelerates from 5.08 to 5.77 m/s (Element A); dotted lines: constant speed of 5.42 m/s (Element B) Figure 18: Total heat transfer for 2.54 mm gauge and all headers ON with various finishing temperatures and mill speeds Figure 19: Comparison of cooling efficiency for various mill speeds, finishing temperatures, and mill accelerations

14

REFERENCES 1) 2) R. M. Guo, "Development, Identification and Application of a Nonlinear Model for Non-symmetric Runout Table Cooling", presented at AISE Annual Conference, 1991, Pittsburgh J. Filipovic, et. al., "Thermal Behavior of a Moving Steel Strip Cooled by an Array of Planar Water Jets", HTD-Vol. 162, Heat Transfer in Metals and Containerless Processing and Manufacturing, ASME 1991, pp 13-23 Izzo, H. F., "A mathematical approach to hot strip mill controlled cooling", Iron and Steel Engineer Year Book, 1972, pp 287-291 Thomas, G., et. al., "A Combined Feedforward-Feedback Computer system for Hot Strip Mill", C.R.M. Metall.Rep. (52), 17-23 May 1978 Seredynski, F. K., "Prediction of Plate Cooling During Rolling-Mill Operation", JISI, 1973, 211, pp 197-203 Yashiro,K., et. al., "Development of Coiling Temperature Control System on Hot Strip Mill", Kawasaki Steel Technical Report, No. 24, Apr. 1991, pp 32-40 J. F. Evans, et. al., "Numerical Modeling of Hot Strip Mill Runout Table Cooling", Iron and Steel Engineer, No.1, Vol. 70, (1993) PP 50 - 55 R. D. Morales, et. al., "Heat Transfer Analysis During Water Spray Cooling of Steel Rods", ISIJ International, Vol. 30, (1990), No.1., pp. 48-57 A. Yousuff, et. al., "Run-out Table Cooling Control", Presented at the "Application of Modern Control in the Metals Industry Seminar -II, Apr. 24-25, 1991, Dearborn, Mich.

3) 4) 5) 6) 7) 8) 9)

10) M. Economopoulos, "Study of Cooling on the Runout Tables of Hot Strip Mills", C. N. R. M., No. 17, 1968, pp 33-42 11) S. J. Chen, A. A. Tseng, F. Han, "Spray and Jet Cooling in Steel Rolling", National Heat Transfer Conf., July 1991, Minneapolis, Minnesota 12) R. Colas and C. M. Sellas, "Computed Temperature Profiles of Hot rolled Plate and Strip During Accelerated Cooling", Proceedings of the International Symposium on Accelerated Cooling of Rolled Steel, Winnipeg, Canada, 1987, Vol. 3, Pergamon Press, London, pp 121-130 13) D. Ellerbrock, et. al., "Characterization of wide-angle spray nozzles for use in accelerated cooling of hot steel bodies", Proceedings of the international symposium on accelerated cooling of rolled steel, Winnipeg, Canada, Aug. 1987, pp 147-157 14) F. Han, S. J. Chen, and C. C. Chang, "The Effect of Surface Motion on Liquid Jet Impingement Heat Transfer", 1991 ASME Winter Annual Conference, Atlanta, Georgia. 15) P. M. Auman, D. K. Griffiths, and D. R. Hill, "Hot Strip Mill Runout Table Temperature Control", Iron and Steel Engineer, Sept., 1967, pp.174-179 16) S. Chatterjee, G. Simonelli, and H. Chizeck, "Parameter Identification of a Nonlinear Coiling Temperature Model for Run Out Table Control at LTV Cleveland Works", AISE, Application of Modern Control in the Metals Industry Seminar -II, Apr. 1991

15

17) J. C. Parks, M. H. Medoff, and J. D. Jesus, "Coiling Temperature control System at Bethlehem Steel Sparrows Point 68-in Hot Strip Mill", AISE, Application of Modern Control in the Metals Industry Seminar -II, Apr. 1991 18) G. Groch, R. Gubermat, and E. Birstein, "Automatic Control of Laminar Flow Cooling in Continuous and Reverse Hot Strip Mill", AISE, Sept., 1990, pp 16-20 19) R. W. Moffat, et. al., "Computer control of hot strip coiling temperature with variable flow laminar spray", Iron and Steel Engineer, Nov. 1985, pp 21-28 20) M. D. Leitholf, and J. R. Dahm, "Model Reference control of runout table cooling at LTV", AISE, Aug. 1989, pp 31-35 21) S. J. Chen, et. al., "Modeling and Analysis of Controlled Cooling for Hot Moving Metal Plates", Symposium of Modelling and Control of Manufacturing Process, ASME PED-Vol.44, 1990 Winter Annual Meeting, ASME, Dallas, pp 465-474 22) Zumbrunnen, D. A., et. al., "A method and apparatus for measuring heat transfer distributions on moving and stationary plates cooled by a planar liquid jet", Experimental Thermal and Fluid Science, Vol. 3, No. 2, pp 202-213, 1990 23) N. Hatta, J. Kokado, and K. Hanasaki, "Numerical Analysis of Cooling Characteristics for Water Bar," Transactions of the Iron and Steel Institute of Japan, Vol. 23, pp 555-564, 1981 24) A. Otomo, S. Yasunaga, and R. Ishida, "Cooling Characteristics of Steel Sheet by Water Film in Hot Strip Mill", J. Iron Steel Inst. Jpn.(JISI), 73, (8), pp 996-1003 25) S. Ishigai, S. Nakanishi, and T. Ochi, "Boiling Heat Transfer for Plane Water Jet Impinging on a Hot Surface", in Sixth International Heat Transfer Conference, Toronto, Canada, August 1978. 26) D. A. Zumbrunnen, et. al., "Convective heat transfer distributions on a plate cooled by planar water jets", Journal of Heat Transfer, Nov. 1989, Vol. 111, pp 889-896 27) R. D. Pehlke, A. Jeyerajan, and H. Wada, "Summary of Thermal Properties for Casting Alloys and Mold, Materials", NSF/MEA-82028, PB 83-211003 28) R. M. Guo, "Numerical solutions for non-symmetric run out table cooling", World Steel Review, Summer issue, 1991. 29) J. Chen, T. Wang, and D. A. Zumbrunnen, "Numerical Analysis of Convective Heat Transfer from a Moving Plate Cooled by an Array of Submerged Planar Jets", HTD-Vol. 162, Heat Transfer in Metals and Containerless Processing and Manufacturing, ASME 1991

16

Das könnte Ihnen auch gefallen