CONVECTION EXPERIMENT LEADER: THOMAS SALERNO JUNE 5, 2006 PARTNERS: GREG ROTHSCHING STEPHEN JOHNSON JEN DIROCCO i ABSTRACT: The convection experiment was separated into two parts. For the first part, we attached a flat aluminum plate inside an air duct. We proceeded to supply a constant 20 watts to the plate while we ran air over the duct at forced velocities of 0, 2.5, 5, and 10 meters per second. Measurements were taken of the steady state temperature of air and of the surface of the plate. The second part of the experiment was exactly like the first, except a finned plate was used for our heated surface. This experiment had three main objectives. The first was to correlate the effect the forced air velocity had on the heat transfer coefficient. The second was to analyze how well our predictive equations were at matching the experimental heat transfer coefficients. Finally, we needed to determine how the addition of fins to the surface aided the heat transfer rate. It was determined from our experiment that the heat transfer coefficient increased linearly with an increase in the square root of air velocity for both the flat and finned plates, over the range of forced convection. From a jump in velocity of 2.5 to 10 m/s, the heat transfer coefficient over the flat plate jumped from 39.4 to 78.1 W/m2 s, while that for the finned plate jumped from 32.1 to 55.8 W/m2 s. However, the limiting value to convection is not zero at zero forced air velocity, but instead a finite value due to natural convection, though the predictive equations were not able to model this effect quite accurately. Also, we saw that the predictive equations yield excellent results in predicting the trend of the heat transfer coefficient, but not the absolute numbers, yielding errors of 80% on the flat plate and 40% on the finned plate. Nevertheless, if one point is taken then we can add a correction coefficient to the predictive equations to increase error to an acceptable one of only 7%. Finally, we determined that the addition of fins increased the heat transfer rate at all velocities, but that its effectiveness diminished as air velocity increased. For the low velocity reading it was able to reduce the surface temperature by 400C, but at the high velocity it only reduced the surface temperature by 180C. The results of this experiment are of great importance for both process and design engineers. This data shows that they can increase the heat transfer coefficient of their system by increasing the flow rate of the heated fluid. Also that predictive equations can be used as long as their geometry matches that of the assumptions for the equations. And if it is close, a simple fudge factor can be added by reading only one data point. It makes the engineer aware that even at zero velocity, a finite and substantial amount of convection will still take place. Finally, the addition of fins will greatly help to increase the heat transfer rate at low fluid velocities, but will not provide a substantial increase at higher fluid velocities. ii TABLE OF CONTENTS ABSTRACT: ____________________________________________________________ ii TABLE OF CONTENTS_____________________________________________________ iii INTRODUCTION:_________________________________________________________ 1 THEORY: ______________________________________________________________ 2 Introduction to Convection: The Convection Problem ____________________________ 2 Figure 1: Convection over flat plate. ______________________________________________ 3 The Velocity Boundary Layer _________________________________________________ 5 Developing the Boundary Layer _____________________________________________________ 5 Figure 1: Flow of fluid over a flat plate. ___________________________________________ 5 Figure 2: Velocity boundary layer forming over a flat plate. __________________________ 6 Describing the Flow Conditions _____________________________________________________ 6 Momentum Equation of the Boundary Layer __________________________________________ 7 The Navier-Stokes Equation _______________________________________________________ 10 The Thermal Boundary Layer _______________________________________________ 11 Developing the Boundary Layer ____________________________________________________ 11 Figure 3: Flow of fluid over a flat plate. __________________________________________ 11 Energy Equation of the Boundary Layer _____________________________________________ 12 The Flat Plate in Parallel Flow _______________________________________________ 13 Solving the Momentum Boundary Layer Equations ____________________________________ 13 Figure 4: Velocity Profile in thermal boundary layer. _______________________________ 14 Solving the Thermal Boundary Layer Equations ______________________________________ 18 Figure 5: Temperature Profile in thermal boundary layer. __________________________ 20 Putting it all Together ____________________________________________________________ 23 Figure 6: Temperature Profile in thermal boundary layer. __________________________ 24 The Flat Plate in Free Convection ____________________________________________ 26 Physical Considerations __________________________________________________________ 26 Figure 7: Boundary Layer development on heated vertical plate. _____________________ 26 The Governing Equations _________________________________________________________ 27 Solving the Governing Equations ___________________________________________________ 29 Finned Surfaces ___________________________________________________________ 31 Physical Situation _______________________________________________________________ 31 Figure 8: Conduction and Convection through parallel fin. __________________________ 31 Establishing Steady State ___________________________________________________ 34 Figure 9: Front view of convection duct. __________________________________________ 35 Figure 10: Side view of convection duct. __________________________________________ 36 Figure 11: Flat Plate and air flow direction. _______________________________________ 36 Figure 12: Finned Plate and air flow direction. ____________________________________ 37 PROCEDURE: __________________________________________________________ 37 RESULTS: _____________________________________________________________ 38 Steady State ______________________________________________________________ 38 Table 1: Steady state temperature values. ________________________________________ 38 Flat Plate Heat Transfer Coefficient __________________________________________ 38 Table 2: Heat transfer coefficients along flat plate. ________________________________ 39 iii Figure 13: Heat transfer coefficients versus velocity for flat plate. ____________________ 40 Table 3: Heat transfer coefficients along flat plate. ________________________________ 41 Figure 14: Heat transfer coefficients versus velocity for flat plate. ____________________ 41 Finned Plate Heat Transfer Coefficient ________________________________________ 42 Table 4: Heat transfer coefficients along finned plate. ______________________________ 42 Figure 15: Heat transfer coefficients versus velocity for finned plate. __________________ 43 Table 5: Heat transfer coefficients along finned plate. ______________________________ 44 Figure 16: Heat transfer coefficients versus velocity for finned plate. __________________ 44 Effectiveness of Fins ________________________________________________________ 44 Figure 17: Temperature Differences recorded for flat and finned plates. ______________ 45 Discussion of Results: __________________________________________________ 46 Effectiveness of Air Velocity on Heat Transfer Coefficient ________________________ 46 Comparison of Experimental results to theoretical ______________________________ 47 Forced Convection ______________________________________________________________ 47 Figure 18: Turbulence over flat plate. ___________________________________________ 48 Figure 19: Turbulence around fins. _____________________________________________ 50 Natural Convection ______________________________________________________________ 51 Effectiveness of Adding Fins to a Flat Plate ____________________________________ 52 Conclusions: __________________________________________________________ 53 Recommendations: _____________________________________________________ 55 References: ___________________________________________________________ 58 APPENDIX:____________________________________________________________ 59 Appendix 1: Sample Calculations ____________________________________________ 59 Run1: Flat Plate Low velocity _____________________________________________________ 59 Table 6: Experimental data obtained for flat plate, at 2.5 meters per second. ___________ 59 Figure 20: Steady state calculation for the first run.________________________________ 60 Table 7: Experimental data obtained for flat plate. ________________________________ 62 Run5: Finned Plate Low velocity ___________________________________________________ 62 Fin Effectiveness for Runs 1 and 5 _________________________________________________ 64 iv INTRODUCTION: In the field of chemical engineering and in every day life itself, convection is one of the most important forms of heat transfer. From running 150,000 lb/hr of hot water through a heat exchanger, to simply turning on a fan to cool down on a hot day, the principles and uses of convection remain the same. Essentially, forced convection is one of the most significant designs of heat exchangers. In order to increase the temperature of their inputs, which allows a faster rate of reaction, or decrease the temperature of waste, or possibly to recover heat that would otherwise be wasted, a majority of processing plants have need of heat exchangers. For example, the pulp and paper industry use heat exchangers to preheat their milling water before it is sent to the separator and, also just as significant, to cool their waste water before it is sent to water treatment. This last principle would permit a plant to recover the heat that would be wasted and channel it for use in other areas of the plant. (CADDET) However, diverse operations will require varying amounts of heat transfers. Thus, process engineers at these plants need to calculate in the area needed for the heat transfer they desire. Such information is only available as a result of convection. Natural convection has many applications. Free convection, strongly, influences the heat transfer from pipes and transmission lines, as well as from electric baseboard refrigeration units to the surrounding air. It is, also, relevant to the environmental sciences, where it manipulates oceanic and atmospheric motions. The most relevant use of natural convection for the chemical engineer, however, is in the cooling of electronic components. Both the performance reliability and life expectancy of electronic equipment are inversely related to the component temperature of the equipment. The relationship between the reliability and the operating temperature of a typical silicon semi-conductor device demonstrates that a reduction in the temperature corresponds to an exponential increase in the reliability and life expectancy of the device. Therefore, controlling the temperature of the device by natural convection (which is a free resource) is of vital importance. (Icoz and Jaluria) Computer engineers, by studying and mastering natural convection, are better able to arrange the area, position, and location of heat sinks to cool their electronic components. 1 To study the effects of convection, I selected to flow air at room temperature on both a flat plate and a finned plate at an elevated temperature for a total of eight trials through a convection chimney. In the first four trials, a flat vertical aluminum plate was placed near the top of the chimney and heated with a constant 20 watts input. Then, air was flowed at 2.5, 5, and 10 meters per second over the plate. Temperature readings of the inlet air and the surface of the plate at steady state were recorded. To find the effect of natural convection, I repeated this setup for an air velocity of zero meters per second. For the remainder of the final four trials, the experiment was then repeated with a finned plate in place of the flat plate. With this data, I was able to correlate how the heat transfer coefficient changed with the air velocity. I was, also, capable of comparing the heat transfer rate of the finned plate to the flat plate at the same air velocity. As a final outcome of the recorded data from my experimentation, I can test predictive equation results with those obtained from the experiment. This experiment had three main objectives. First, to study the effect the air velocity had on the heat transfer coefficient; second, to examine the accuracy of current predictive equations and finally, to find the effectiveness of the fins at different air velocities. With this information, the principles of designing heat exchangers take form. THEORY: Introduction to Convection: The Convection Problem Heat transfer due to convection involves the transfer of energy between a fluid at one temperature moving over a solid surface at another temperature. Consider the situation shown in Figure 1. 2 T8 A8 u8 Ts A s L Figure 1: Convection over flat plate. Here a fluid, namely air, of velocity u (meters/sec) and temperature T∞ (°C) flows over a flat aluminum surface of length L (meters) and area As (meters2). The temperature of this surface is assumed to be uniform at Ts (°C) such that Ts > T∞. This will result in a temperature gradient forming from the surface and extending into the fluid. We know from Newton’s Law of Cooling that heat transfer at the surface will occur via a rate that is proportional to the difference between the surface temperature Ts and the temperature of the fluid T∞. Thus, we can write the local heat flux (heat transferred per unit area) as (1) q h(Ts T ) A Where q (Watts) is the heat transferred per unit time, A (meters2) is the surface area available for heat transfer, and h (Watts/m C) is the constant of proportionality known as the local heat transfer coefficient. Note that both q and h are referred to as local. This is because flow conditions will vary from point to point along the plate causing both q and h to vary with position x (meters). The total heat transfer rate Q may be obtained by integrating the local heat flux over the entire surface. That is, (2) Q q A dA s As or from Equation 1 we can write, 3 (3) Q q A dA h(T s As s T )dAs (Ts T ) hdAs As As We can then define an average heat transfer coefficient h (W/m C) as, hdA s (4) h As dAs 1 As hdA . s As As Thus, the total heat transfer rate may also be expressed as (5) Q hAs (Ts T ) . Note, that for our flat plate, we are considering flow in only one direction, namely the x direction. Thus, we may rewrite Equation 4 as L (6) h 1 hdx . L 0 Equation 5 illustrates two very fundamental uses. First, with this equation, we have available a method to experimentally measure our average heat transfer coefficient over a flat plate. In short, if one has all the variables (measure of heat flux being convected from the plate; the surface area of the plate, the temperature of the flowing air, and the steady state temperature of the surface of the plate ) needed in the equation then, one is able to obtain the average heat transfer coefficient for the flat plate. Secondly, to find the total heat transferred from a heated plate to a flowing fluid, we only need to know the average heat transfer coefficient, surface area of the plate and the temperature of the surface and of the bulk fluid. However, in order to arrive at the average heat transfer coefficient, we must, first, know how the local heat transfer coefficient varies with x direction along the plate. This is dependent upon numerous fluid properties such as viscosity, density, thermal conductivity, specific heat etcetera as well 4 as the flow conditions along the plate. This multitude of independent variables in determining the average heat transfer coefficient has been termed the problem of convection. One way to determine all of these variables is to consider the concept of boundary layers. The Velocity Boundary Layer Developing the Boundary Layer To develop the concept of boundary layers, consider, again, the flow situation represented by Figure 1 below. T8 A8 u8 Ts A s L Figure 1: Flow of fluid over a flat plate. When fluid particles make direct contact with the surface, we assume their velocity is reduced to zero due to viscous action between the surface and the fluid. These particles will then act to retard the motion of particles in the adjoining fluid layer to a lower velocity, which will then retard the motion of the next fluid layer and so on. The retardation of fluid motion is due to shear stresses (N/m2) acting between parallel fluid layers. This retarding of motion will continue until we reach a distance where the velocity of the fluid particles flowing over the plate are unaffected by the plate’s presence and continue at a velocity of u . It is this vertical distance from the plate to the first particle whose velocity is unaffected is (meters) that is known as the boundary layer thickness. Accordingly, we can develop a boundary layer velocity profile which shows the manner in which u varies with the y direction throughout the boundary layer. This is represented below in Figure 2. 5 u8 T8 A8 u8 TS AS L Figure 2: Velocity boundary layer forming over a flat plate. This boundary layer will form whenever a fluid flows over a flat surface. Its shape will depend on how the shear stress varies with shear strain u in the fluid as well as the flow y conditions we are experiencing. Assuming, we have a Newtonian fluid, we can evaluate the shear stress as, (7) u y where (kg m/sec) is the viscosity of the fluid. As for the flow conditions, there are typically three patterns seen. Describing the Flow Conditions First, in laminar flow, the motion of air particles is very orderly with all particles at a position y moving in a straight line parallel to the pipe walls. Secondly, in turbulent flow, the particles move around and have different velocity components in all directions. Thirdly, a transition flow has mixed properties of the previous two. From observation, it was hypothesized that fluid travels in a laminar motion when it is moving slowly and converts to turbulent motion when it has a faster velocity. But this information was not very useful in that it only gave us qualitative definitions of our flow. As a result, in the 1880’s, Osbourne Reynolds designed an experiment which involved inserting dye into the center of a water stream which was flowing through a glass tube and where the 6 velocity was controlled by a valve. After many experiments, he developed the following equation, (8) Re x ux where Re is called the Reynolds number, ρ is density (kg/m3), and x is the distance along the plate in the x direction. He found that the dye made a straight line and stayed in the center when Re < 5 105 (laminar), and that the dye formed fishhooks when Re > 107 (turbulent). Later this term was given a more scientific understanding, namely (9) Re inertial forces . viscous forces Thus, when inertial forces such as pressure and mass flow overcome the viscous forces, we have a large Re and our flow is very turbulent, however, if the viscous forces are larger, than all of the particles stick together and stay in line and we arrive at laminar flow. Momentum Equation of the Boundary Layer In order to study the motion of particles, we need a method to model the interaction and motion of any fluid element in the boundary layer. This data can be obtained by performing a momentum balance on the fluid particles and developing what is known as an equation of motion for them. Consider an arbitrary volume element. Momentum can be transported into and out of this volume element by three different mechanisms. The first is through convection, the bulk flow of fluid across the surface. The second is via forces that act on the surface of the volume element and the third is by body forces which act on the entire volume of the element in order to change its momentum. We are able to arrange these three mechanisms to fit a verbal differential equation shown below. Rate of Net rate of Net rate of Net rate of 7 (10) change of momentum = momentum + momentum convected into V + momentum creation by creation by surface forces body forces Equation 10 is the verbal representation of the momentum balance on this fluid element. The next step is to put this equation into mathematical terms. The rate of momentum change caused by convection is determined by the rate at which mass will enter and leave this fluid element and the velocity of this fluid. The local volumetric flow rate of fluid across a surface element dA is the velocity of fluid multiplied by the area, or in other words, u·dA. Thus, the rate of mass transfer in and out of this element is given by ρ(u·dA). Multiplying this by the velocity determines our first term of the equation. Note, that this equation is negative because it assumes the momentum is being convected into the surface element. If momentum is in fact leaving the surface element in a larger amount than the dot product, then, this will ensure that the term turns out positive. (11) Rate of momentum convected u u d A The rate of momentum change created by surface forces is caused by molecular motion and interactions within the fluid itself. This rate of molecular diffusion can be described by the shear stress on the volume element multiplied by the area on which this stress acts. These shear stresses are acting to compress this volume element and, thus, this term must also be negative. (12) Rate of momentum created d A by surface forces Body forces acting on our volume element may be gravitational, surface, electrical, or magnetic. However, in the present situation, the most common forces are only gravitational and surface, and will be the only ones considered here. The gravitational force is, simply, the acceleration due to gravity pulling down on the mass of the element, where the surface force is the external pressure acting to compress our volume element. 8 (13) Rate of momentum created Pd A g dV by body forces Finally, we can write the rate of accumulation of momentum into and out of this volume element as, (14) Rate of accumulation of momentum u dV t Thus, the momentum balance can now be written mathematically as: (15) t u dV uu d A d A Pd A gdV C .V . C .S . C .S . C .S . C .V . Mathematically, C.V. signifies an integral across the control volume of our fluid element and C.S. represents an integral across the surface of our fluid element. However as it stands, this equation cannot be solved easily. In order to simplify this equation, we would need to organize our data into the same units and transform the integrals into differentials which can be more easily applied to a given situation. Fortunately, we have from Gauss’s theorem which is the ability to convert each integral over the control surface to an integral over the control volume thereby standardizing the integrals. This is achieved by using the gradient vector, as shown below: (16) u d A udV C .S . C .V . Thus, we can now write Equation 15 as: (17) t u dV uudV dV PdV pgdV C .V . C .V . C .V . C .V . C .V . 9 Now, Equation 17 illustrates that all terms have a triple integral over the control volume. Provided that no math teachers are looking, we can cancel out all the integrals and arrive at: (18) dV P p g t Equation 18 is what is generally referred to as the equation of motion for a fluid. Just to be certain nothing is lost in the math, we will define the physical significance of each term. The term on the left side of the equal sign is the rate at which we accumulate momentum in our fluid element. On the right side of the equation, the first term describes the change in momentum due to convection via fluid flow. The second term accounts for change in momentum due to molecular diffusion which is based off the viscosity of the fluid and the shear stresses placed on it. The third term describes the change of momentum that is caused by a pressure drop due to the friction experienced by flowing along a plate. Finally, the last term on the right takes into account the gravitational force on our fluid element. The Navier-Stokes Equation Inserting Equation 7 into Equation 18, we have developed what is known as the NavierStokes equation where we assume that we have a Newtonian fluid, a constant density, and a constant viscosity. This equation is shown below for flow in the x direction in a Cartesian coordinate system: (19) v x P 2 v v x v x g x t x However, we could greatly simplify Equation 19 by factoring in a few more assumptions. First, we will assume that flow is only in the x direction and that momentum is changing only in the y direction. Secondly, we will again assume that the pressure drop for our external flow is negligible. Applying these assumptions to Equation 19 and dividing through by and noting that is the molecular diffusity of the fluid, we arrive at 10 the working momentum equation of the boundary layer under laminar flow conditions, namely, (20) u u 2u . u x y y 2 The Thermal Boundary Layer Developing the Boundary Layer The development of the thermal boundary layer is similar to that of the velocity boundary layer. Consider the flow situation represented by Figure 3, shown below. T8 T8 A8 u8 TS AS L Figure 3: Flow of fluid over a flat plate. At the leading edge, the temperature profile of the fluid is uniform. However, once the fluid makes contact with the plate the particles at the surface will achieve thermal equilibrium at the plate’s surface temperature. These particles will, then, exchange energy with the layer of fluid particles above them, who will exchange energy with the fluid particles above them. This will form a temperature gradient in the fluid from the surface at a temperature of Ts to a distance t (meters) where the temperature is that of the bulk fluid T∞. Here t is known as the thermal boundary layer thickness and is the region of the fluid in which the temperature gradients exist. 11 Energy Equation of the Boundary Layer Just as we developed the momentum equation of the boundary layer, we will develop the energy equation of the boundary layer by performing an energy balance on the fluid particles inside the boundary layer. Energy is carried in and out of any arbitrary volume element by three mechanisms. Convection is the first mechanism by the bulk flow of fluid across the surface. The second is through conduction, the energy transfer from molecular vibrations of particles adjacent to the control element. With the third being viscous work where the body forces act on the entire volume in order to change its momentum. By combining all three mechanisms, we arrive at a verbal energy equation shown below in Equation 20. (21) Energy convected Energy convected Heat conducted Net viscous work in at x in at y in at y done on element Energy convected Energy convected Heat conducted out at x dx out at y dy out at y dy Note: that energy being convected into a control volume element is calculated by the mass flow rate entering multiplied by the enthalpy of that mass. In mathematical terms we transcribe it as, (22) Energy convected C pu x positionTx position dy Also, the energy being carried in by conduction is, according to Fourier’s Law, equal to the conduction heat transfer coefficient multiplied by the temperature gradient at this position. Hence, in mathematical terms we can write. (23) T Energy conducted kdx y y position Finally, the viscous work being done on the element is equal to the viscous-shear force acting on the element ( unit time ( u dx ) multiplied by the distance through which it moves per y u dy ). So, we may write this mathematically as, y 12 2 (24) u Viscous work dxdy y Therefore, we can plug Equations 22,23, and 24 into Equation 21 to yield the differential energy balance, T u u T 2T C p u T dxdy k dxdy dxdy 2 y y x y y x 2 (25) u Then using the continuity relation 0 and dividing through by C p , we x y arrive at, (26) T T 2T u u x y y 2 C y p 2 which is the energy equation of the boundary layer under laminar flow conditions. In this equation k , is known as the thermal diffusity of the fluid. However, this can be C p further simplified by realizing that the second term on the right side of Equation 26 is only of importance for highly viscous fluids and is therefore negligible for air in our operating temperatures. This leaves us with our working energy equation for laminar boundary layers, namely, (27) T T 2T u . x y y 2 The Flat Plate in Parallel Flow Solving the Momentum Boundary Layer Equations Consider the flat plate situation shown below in Figure 4. 13 A H A u8 dy δ dx Figure 4: Velocity Profile in thermal boundary layer. In order to solve for a velocity profile and the boundary layer thickness along the plate, we must perform a momentum balance and a force balance on this arbitrary element. Starting with the momentum balance we have, (28) Net Momentum Flow Momentum carried out Momentum carried in Since we know that momentum in our situation is only carried in and away by mass, we can make use of the equation Momentum mass flow * velocity and write for the momentum entering through the left side as H (29) u dy 2 0 and the momentum through the right side is H d 2 u dy u dy dx . 0 dx 0 H (30) 2 14 The momentum flow through the bottom must be zero since there is a solid plate there, and we are left with only the momentum through the top. This equation is complicated. First, we need to know the mass flow entering the top. This can be found from the difference between the mass exiting through the right and the mass entering through the left as there is no creation or destruction of mass in this volume element. Thus, the mass flow through the top is, (31) H H H H d d u dy dx u dy u dy u dy dx . dx 0 dx 0 0 0 This mass must carry with it a momentum equal to its mass flow rate times its velocity. Since this area is above the boundary layer, we know that its velocity must be that of the free stream. Therefore the momentum through the top is, (32) H d u u dy dx . dx 0 Combining Equations 32, 30, and 29 into Equation 28 yields us with our net momentum flow out of our element as, (33) Net Momentum Flow H H d d 2 u dy dx u udy dx . dx 0 dx 0 Then using the product rule from calculus, we can rewrite the second term in this equation as (34) u H H H du d d udy dx uu dy dx udy dx . dx 0 dx 0 dx 0 Accordingly, we can rewrite Equation 33 in more useful terms as 15 Net Momentum Flow (36) H H d H du d 2 u dy dx uu dy dx udy dx dx 0 dx 0 dx 0 H du H d (u u )udy udy dx 0 dx 0 Next, we need to do a force balance on this same volume element. The force at the left and right sides are due to pressure force. The force at the left side is pH and that at the dp right side is p dx H . Due to shear stress from the wall, the force at the bottom dx can be written as w dx dx u y , which is the definition of viscosity. That leaves y 0 only the force at the top. However since this section is outside our boundary layer, we know there are now velocity gradients here and thus no force. Collecting terms, we arrive at the net force balance as Net Force Force out Force in (37) dp Net Force pH p dx H w dx . dx dp Net Force w dx H dx dx Then from Newton’s Law, the net force on an object is equal to its net increase in momentum, we can set Equations 36 and 37 equal to each other to obtain. (38) H du H d dp w H (u u )udy udy dx 0 dx dx 0 For our system, we will assume that pressure across the plate is relatively constant allowing us to cross out the second term on the left hand side of the equation. As a direct result of the constant pressure assumption, we know the free stream velocity is not a function of x, consequently, we can cross out the last term on the right hand side of the 16 equation. Finally from our definition of the boundary layer, we know that u=u∞ for y> δ, thus we can change the limit on the integral to δ instead of H. This leaves us with, (39) u w y y 0 d (u u )udy dx 0 where we have brought in our definition of viscosity to rewrite the shear stress as a function of the velocity gradient. The next step is to find an equation for the velocity profile for use in Equation 39. Using an approximate development, we can assume that this relation will be in the form of a simple polynomial. To calculate approximately how many terms we need, let us list the conditions that must be satisfied: (a) the velocity at the surface must be zero due to viscous forces; and since our definition of our boundary layer is the only region in which velocity gradients exist we arrive at: (b) the velocity at the top of the boundary layer must be equal to the free stream velocity; and (c) the velocity gradient must be zero. For our final condition, since we are at a constant pressure situation, we can solve Equation 20 with our other conditions to obtain (d) the velocity gradient must stop changing at the plate surface. We can express these mathematically as, (a) u =0 at y=0 (b) u = u at y= (c) u 0 at y= y (d) 2u 0 at y=0. y 2 The simplest polynomial that could satisfy all four conditions is, (40) u C1 C2 y C3 y 2 C4 y 3 . 17 Applying the four conditions (a) through (d), we arrive at an expression for the velocity as a function of y position in the boundary layer, (41) u 3 y 1 y u 2 2 3 for Re 5 1010 . We can now plug Equation 40 into Equation 39, perform the integration and solve for , x d 39 3 u u2 dx 280 2 140 140 d dx dx 13 u 13 u (42) 2 2 140 x Const 13 u @ x 0, 0 Const 0 x 4.64 Re 1 2 x 2 2 140 x 13 u for Re 5 1010 which will then tell us the boundary layer thickness as a function of x along the plate. Solving the Thermal Boundary Layer Equations In order to find the temperature gradient in the boundary layer, we must solve the energy equation of the thermal boundary layer, Equation (27). The conditions that this equation must satisfy are: (a) the temperature at the surface of the plate must equal the surface temperature because we have zero velocity at this point and the volume elements here must be in thermal equilibrium with the plate; (b) the temperature gradient should be zero at the top of the thermal boundary layer. This is a direct result of our definition that the thermal boundary layer is the area in which we experience temperature gradients, thus the temperature gradients must stop as we leave the boundary layer; (c) the temperature at the top of the boundary layer must equal the free stream temperature, this is also a direct result of our definition of the thermal boundary layer, since we are outside the area of 18 temperature gradients, we are outside the area which is affected by the plate, thus our temperature must equal the free stream temperature we originally had; (d) finally the rate of change of the temperature gradient must also be zero at the wall because the velocities are zero at the wall thus there is no driving force to change the temperature gradient. We can express these conditions mathematically as (a) T=TS at y=0 (b) T =0 at y=δT y (c) T=T∞ at y= δT (d) 2T =0 at y=0. y 2 We now assume that the temperature profiles at various x positions must all have the same functional dependence on the y position. Consequently, the simplest method to solve Equation (27) with the four conditions (a) through (d) is to fit them to a third degree polynomial with arbitrary constants, namely, (43) T Ts C1 C2 y C3 y 2 C4 y 3 T Ts Then after applying the four conditions, we are able to arrive at (44) 3 y T T 2 T 1 y 2 T 3 for Re 5 1010 Now, we have found an equation for the temperature profile of the thermal boundary layer. However, we now need to find a relation for the thermal boundary layer thickness. This can be achieved by an integral analysis of the energy equation of the boundary layer shown again in Figure 5 below. 19 A H A u8 T8 δT δ dx dq w kdx dT dy w Figure 5: Temperature Profile in thermal boundary layer. Performing an energy balance yields (45) Energy convected in Viscous work within Heat transfer at wall Energy convected out H The energy convected into the left side is, C p uTdy ; the energy convected out the right 0 H d H side is, C p uTdy C p uTdy dx ; the energy due to mass flow through the 0 dx 0 top is, C pT H d udy dx ; the net viscous work done within this element is, dx 0 H du 2 T . Combining all dy dx ; and the heat transfer at the wall is dqw kdx y w dy 0 of the above relations into Equation 45 yields, (46) 2 H d H du T dy (T T )udy dx 0 y C p 0 dy w 20 However, the net viscous work term is negligible unless we have a highly viscous fluid, which air is not; therefore we will ignore this term and arrive at, (47) H d T (T T )udy dx 0 y w This is the energy boundary layer equation for our laminar flow over a flat plate. The next step in calculating the thermal boundary layer thickness is to insert our temperature profile Equation 44 and the velocity distribution Equation 41 into Equation 47, which will then look like this, (48) H d H d (T T )udy ( )udy dx 0 dx 0 3 3 H d 3 y 1 y 3 y 1 y u 1 dy dx 0 2 T 2 T 2 2 T y w 3 2T where we have defined = T – TS. Allow us to assume δT is smaller than δ, which is the case for air and most fluids, we can change the integrals to sum up to δ instead of H. Performing the necessary integration we arrive at, (49) 2 4 d 3 T 3 T 3 u dx 20 280 2 T and because T <1 we can assume T is negligibly small compared to T and we 4 2 can write 21 (50) 2 3 d 3 u T 20 dx 2 T performing the differentiation gives (51) t 2d 2 2 t 1 u dx 10 3 t d dx But from the development shown in Equation 42 we know, 2 2 140 x so that we can 13 u write, t d t t 4x dx 3 (52) 2 t d t but noting that dx 2 13 , 14 1 d 3 t we realize that Equation 52 is a first order 3 dx linear differential equation whose solution is 3 13 t 4 Cx 14 3 (53) For our situation, we will assume that the boundary layer starts forming at x=0 but that the heat of the plate doesn’t start until x=x0. These will give us our two boundary conditions, namely, 22 (a) t 0 @ x x0 (b) t 0 @ x x0 We can solve these simultaneously with Equation 53 to yield (54a) t 1/ 3 x0 3/ 4 1 1/ 3 1.026 Pr 1 x , or if the heat of the plate starts immediately at x=0, (54b) t 1 1/ 3 1.026 Pr for Re 5 1010 . In these equations we noticed a ratio that has been in numerous equations throughout this development. This ratio is CP . Looking more deeply into the properties of this k ratio we see it is a ratio of the kinematic viscosity. A ratio of this kind of viscosity conveys information about the rate at which momentum diffuses through a material, and the thermal diffusivity, which conveys how heat diffuses through a material, each by molecular vibrations. It is through this ratio that we discover the relative magnitudes of momentum and thermal diffusivity in the material, which is the information needed to determine the relative sizes of the momentum and thermal boundary layers. This link between the velocity field and the temperature field is so important that it was named, as is shown in Equations 54, Pr, the Prandtl number. Putting it all Together Visualize, again, the situation where we have a fluid flowing over a hot plate, shown below in Figure 6. 23 T8 T8 A8 u8 TS AS L Figure 6: Temperature Profile in thermal boundary layer. The temperature of the surface is at Ts, the temperature of the fluid is at T∞, and the boundary layer thickness is at a height of δT. At the wall we have zero velocity due to viscous forces, thus the heat transfer from the wall into the boundary layer must be done through conduction. Therefore, our heat flux into the boundary layer must be (55) Q T ks As y surface We can now combine this with Equation 5 and Equation 6 to obtain, k (56) h T y wall Ts T 3 k . 2 t We can now plug in Equations 54 and Equation 42 to yield, (57) hx .332k Pr 1/ 3 Re x1/ 2 x x0 3/ 4 1 x 1/ 3 However, from our development, we realize that we can use this equation for all similar geometries. Hence, we wish to nondimensionalize this equation such that it will be 24 applicable to all flat plates that we will encounter. We will divide both sides by x / k which will give us, (58) x0 3/ 4 hx x 1/ 3 1/ 2 Nu x .332 Pr Re x 1 k x 1/ 3 It is at this point that we have defined another useful dimensionless parameter known as the Nusselt number. As is stated in the above paragraph this makes for an important parameter. Through the development of a flat plate, we expect that we can create a single correlation for a given similar geometry and then use dimensionless parameters such that this correlation can be used for all fluids and all absolute dimensions of the relative geometry. The Nu, Pr, and Re all provide us with this ability. In fact, from this development, we can now state with a fair amount of certainty that h is a function of Nu which is a function of Re and Pr. Thus for all other geometries, we can run numerous experiments with different fluids and create correlations for h based on different manipulations of the these three dimensionless parameters. Since most of the geometries are very difficult to solve analytically, this is the exact procedure we use to obtain the correlations for a number of turbulent flows in different geometries. Back to the analysis, we want to extend our relation for the heat transfer coefficient for the flat plate in forced flow to include one more step. We want to have the average heat transfer coefficient over the entire plate such that we can use an average equation for the entire plate instead of integral analysis every time. From the definition of average, we can calculate L h dx x (59) h 0 L dx 2hL 0 Thus we can plug Equation 59 into Equation 58 and obtain the working relationship for forced laminar flow over a flat plate as, 25 (60a) Nu hL .664 Pr1/ 3 Re x1/ 2 k 1/ 2 L .332 Pr1/ 3 k u x (60b) h 0 L for Re 5 1010 3/ 4 1 x0 1 x x 1/ 3 dx for Re 5 1010 The analysis to arrive at this equation assumed that all physical properties were constant. If this is not the case, it is suggested by many textbooks to use properties at the film temperature, (61) Tf Tw T . 2 The Flat Plate in Free Convection Physical Considerations In our previous discussion, we examined the situation of a flat plate and a fluid with a forced velocity flowing over it causing forced convection. However, convection can also result when there are density gradients caused by temperature gradients producing the net effect of buoyancy forces on the molecules present due to the gravitational field. Consider the situation represented in Figure 7 ρ = fucn(y) T8 ρ8 U = func(y) Ts Figure 7: Boundary Layer development on heated vertical plate. 26 In this situation, a heated vertical plate at Ts is surrounded by air at T∞. The fluid close to the plate is at a higher temperature and less dense than fluid far removed. Therefore, buoyancy forces induce a free convection boundary layer in which the fluid at a higher temperature rises vertically, entraining fluid from the fluid far removed. The resulting velocity distribution, shown in Figure 7, is unlike that we have seen before and thus requires a new development. The Governing Equations In order to analyze this situation, we again need the momentum and energy equations that we derived from the related conservation principles. Thus, for momentum, we return to Equation 19, (19) v x P 2 v v x v x g x t x We can, again, simplify this equation by assuming that flow is only in the x direction and that momentum is changing only in the y direction. However, we cannot ignore the pressure drop term, as its value is no longer negligible nor can we ignore the body force term that gravity imparts on our system. Applying these to Equation 19 and dividing through by , noting that (62) is the molecular diffusity of the fluid, we arrive at, u u 1 p 2u u g 2 . x y x y We can rewrite this equation by noting that the x pressure gradient in the boundary layer must be equal to the pressure gradient in the surrounding fluid, but in the surrounding fluid u=0. Applying this we can solve Equation 62 for the surrounding fluid to find, (63) p g x Substituting Equation 63 into Equation 62 yields, 27 (64) u u 2u u g 2 . x y y Note: the new term in this equation represents the buoyancy force present in our system due to the density gradient. We can make this equation more functional if we assume that density variations are caused solely by temperature variations, in which case we may take advantage of the fluid property known as the volumetric thermal expansion coefficient. This property measures the change in density with a change in temperature at constant pressure. (65) 1 T p However, we will further simplify Equation 65 with the Boussinesq approximation, which allows us to approximate the derivative so that we arrive at, (66) 1 1 T T T which can be solved for and then directly substituted into Equation 64 to yield the working momentum equation for free laminar flow along a vertical plate, (67) u u u 2u g (T T ) 2 . x y y Again note, that the buoyancy force as we are representing it only has an effect on our momentum equations, thus we can rewrite our working expressions for the mass and energy equations exactly how they were for forced convection. This provides us with our three working equations from which we can solve our free laminar boundary layer, namely, 28 (68) u 0 x y (69) u (70) T T 2T u x y y 2 . u u 2u g (T T ) 2 x y y Solving the Governing Equations If we nondimensionalize the equations, we realize that we can obtain much more practical expressions for our correlations. Hence, we define the following variables x T T x y u , y , u , , T , where L is a characteristic length and u0 L L u0 u0 Ts T is an arbitrary reference velocity. Applying these substitutions into Equations 69 and 70 yield the following nondimensionalized free laminar boundary layer equations (71) g (Ts T ) L u 1 2u u u T x y u0 2 Re L y2 (72) u T 1 2T T x y Re L Pr y 2 where we have a new dimensionless parameter as a result of the buoyancy force, g (Ts T ) L . However, in its present form it is problematic because it requires some u0 2 2 u L reference velocity. Thus, we will multiply this parameter by Re L 2 o and obtain what is known as the Grashof number, g (Ts T ) L uo L g (Ts T ) L3 . GrL u0 2 2 2 (73) 29 The Grashof number is the free convection equivalent to the Reynolds number in forced convection. Where Reynolds was the ratio of inertial to viscous forces, Grashof is the ratio of the buoyancy force to the viscous force acting on the fluid. In order to solve Equations 71 and 72 for our situation, we can take two directions. The first involves us creating differential elements and balancing momentum and forces, deriving expressions for the velocity and temperature distributions, and finally solving for a functional relationship of the boundary layer thickness. Such a route is quite tedious and provides below satisfactory results for its “exact” correlation. The second method we can take is to recognize that in forced convection, the Nusselt number was a function of the dimensionless parameters present in forced convection, Pr and Re. In consequence, we can assume that once again we are going to experience the Nusselt number to be a function of the dimensionless parameters present and write, (74) Nu L function(Pr,GrL ) . Then, we simply need to run a multitude of experiments to determine the correct correlation for our system. This is the method utilized by Churchill and Chu, according to Welty, Wicks, Wilson, and Rorrer. In their process, they defined a new dimensionless parameter which is simply a manipulation of our current dimensionless parameters. This new parameter is called the Rayleigh number and is defined as, (75) RaL GrL Pr g (Ts T ) L3 . They then anticipated the functional relationship of Equation 74 would be (76) Nu L hL CRa L n . k Churchill and Chu proceeded to run multiple experiments in the range of RaL 109 and found the working relationship for laminar flow to be, 30 (77) Nu L .68 .670 Ra1/L 4 .492 9 /16 1 Pr 4/9 . Finned Surfaces Physical Situation It is a frequent practice to add extended surfaces to a flat plate to increase the rate of heat transfer away from the surface. As shown in Figure 8 below, we have added rectangular fins to the plate. dqconvection = hf *P dx * (Tfin @ x - T8 ) P t z dx L Figure 8: Conduction and Convection through parallel fin. Heat generated in the surface is being conducted through the fin and then convected away to the surrounding fluid. The temperature at the base of the fin is TS and the temperature of the surrounding fluid is T∞. To begin our analysis of this situation, we must first conduct an energy balance around our fin.as calculated below 31 Energy conducted in Energy conducted out Energy lost by convection kA (78) dT dT kA dx dx hPdx(T T ) x dx dT d 2T dT kA 2 dx hPdx(T T ) dx dx dx 2 d T hP 0 dx 2 kA kA where T T . In order to solve the above differential equation, we need two boundary conditions. To simplify the math, we will assume the tip of the fin is insulated, which basically means that there is no driving force for heat transfer at the tip which is mostly the case. Thus, we can now label our two conditions as follows: (a) the temperature of the fin at its base is equal to the temperature of the surface; and (b) there is no temperature gradient at the tip (insulated). Mathematically we can write these conditions as, (a) 0 @ x 0 (b) d 0 @ dx xL Solving the above second order linear differential Equation 78 with the two boundary conditions a and b, we find the temperature distribution along the fin as, (79) cosh(m( L x)) 0 cosh( mL) where we have defined a new constant, m 2 hP , to make the math easierr to follow. kA Noting that all of the heat lost by the fin via convection must be conducted into the base of the fin by conduction, we may write an expression for the total heat lost by the fin as, 32 q kA (80) dT dx x 0 1 1 q kA 0 m 2 mL 1 e 2 mL 1 e q hPkA 0 tanh(mL) To make this relation more applicable, we would like to combine the heat convected from the plate and the heat convected from the fin into one equation. In order to do this, we define the fin efficiency as follows, (81) Fin efficiency actual heat transferred f heat which would be transferred if entire fin area were at base temperature For our case ,we can determine this mathematically as (82) f hPkA 0 tanh(mL) tanh(mL) . hPL 0 mL For this reason, we can describe convection from a flat plate with fins as, (83) q h0 A0 h f Af f T T s where h0 A0 are for the flat plate without fins and h f Af f from our equations will be the heat convected from the fins. Consequently, all we need to do is measure the areas of the flat plate and the fins and the temperature of the surface and the surrounding air, and the total heat produced within the plate to find the heat transfer coefficients of the plates and of the fins. Of particular note here, we may also find an estimated heat transfer coefficient of the fins by using Equation 60, which we developed for flat plates. This is a good approximation 33 because the fins are in essence flat plates but with less uniform temperatures. However, because the conductivity of the plate is very high, we can assume that the temperature is fairly uniform. Establishing Steady State There were a total of eight runs used for this experiment. Six of these were for forced convection with the other two for natural convection. As for the latter, steady state values were obtained by allowing the system to run for an hour and a half. Measurements were then taken for each variable. Because of the extended run time, these readings were assumed to be steady state values. For the six forced convection runs, measurements were taken every 5 minutes for an hour, or until the readings started leveling out. This data was then inserted into Newton’s Law of Cooling model to determine the steady state values. From Newton’s Law of Cooling, we know the rate of change in temperature is proportional to the driving force. In this case, a constant times the difference between the steady state temperature and the current temperature, or in mathematical terms, (84) dT k T T dt Solving this separable differential equation we arrive at (85) T T exp Kt , K Constant Thus, we can graph Temperature versus and guess values of K to give us a good correlation. The y-intercept of this graph will be the steady state temperature the system is trying to attain. 34 EQUIPMENT: For these experiments, the medium of heat transfer was air, which was flowed through a duct with the use of a pump. The duct (chimney) has a cross sectional area of 120mm by 70mm and stood about 1.5m tall. Air velocity was controlled by the variac controller on the pump and could be further adjusted by a slide cover in front of the chimney. Measurements of air velocity were made by a portable analog anemometer mounted on the duct. Readings were taken in meters per second. A 20 watt power control circuit manufactured by Armfield Technical Education Co., HTG-B Serial Number 3681-3, which had a direct reading digital watt meter imbedded in it, heated the surface. Temperature measurements were taken to a resolution of .10C through the use of thermistor probes and a digital temperature indicator which read in 0C which is located on the control circuit. All equipment is shown below in Figures 9 and 10. Chimney Viewing Window Anemo meter Figure 9: Front view of convection duct. 35 Boundary Layer Profile Measurement Heated Surface Inlet Air Measurement Temperature Probe 20 watts Pump and Slide Cover Power Supply Figure 10: Side view of convection duct. There are two surfaces that will be used in this experiment both are made from an aluminum alloy with an estimated conductivity of 166.5 Watts . The first surface is a flat m 0C plate, with a length of 100mm and a width of 110mm. Subsequently, the second surface is a finned plate, which has 9 fins that are 4mm thick at the base, 2mm thick at the top, 100mm long (in direction of air flow) and 68mm in depth. These 9 fins have a horizontal spacing of 12-13mm across a flat plate identical to the first surface. Pictures of these surfaces are shown below in Figures 11 and 12. T8 A8 u8 Figure 11: Flat Plate and air flow direction. 36 T8 A8 u8 Figure 12: Finned Plate and air flow direction. PROCEDURE: 1. Insert the flat plate in the proper position in the chimney. Plug in the power supply to the flat plate. Attach the temperature sensor into the flat plate. 2. Insert the Temperature probe into the air inlet hole. 3. Turn on the anemometer. Turn on the pump and adjust the variac controller and the slide cover to achieve an air velocity of 2.5m/sec. 4. Turn on the power source. Adjust power output to 20 Watts. 5. Start taking readings every 5 minutes until the system reaches steady state, or after an hour of measurement time, which ever comesfirst. 6. Move the temperature probe into each of the three profile holes near the middle of the plate. Allow the probe 5 to 10 minutes to reach steady state and record the reading. 7. Document all data. 8. Repeat steps 2 through 7 for air velocities of 5 and 10 m/sec. Then, repeat steps 2 through 7 once more with the pump turned off to record a measurement for natural convection. Note this last run may take up to two hours to reach steady state. Readings do not have to be taken for every 5 minutes as long as you leave enough time for the system to reach steady state. 9. Turn the power source off. Unplug the flat plate power source and temperature sensor. Remove the flat plate. Insert the parallel fins. Plug the power source and the temperature sensor into the parallel fins. 37 10. Repeat steps 2 through 8 for the parallel fin setup. 11. Turn the power supply off. 12. Clean up the lab. RESULTS: Steady State For the six forced convection runs, we arranged the data to a graph of temperature versus exp(k t ) . The data for this can be seen in Appendix 2. The two natural convection runs had actual steady state measurements taken during their analysis, therefore no manipulation of data was required to obtain their steady state values. The steady state values of all runs are shown in Table 1 below. Table 1: Steady State Temperatures Flat Plate: Air Velocity(m/sec) Air Inlet Temp (°C) Plate Surface Temp (°C) 0.0 25.5 93.8 2.5 25.7 71.8 5.0 24.9 58.8 10.0 25.0 47.1 Finned Plate: Air Velocity(m/sec) Air Inlet Temp (°C) Plate Surface Temp (°C) 0.0 25.0 47.1 2.5 25.0 31.1 5.0 25.0 29.9 10.0 25.0 28.9 Table 1: Steady state temperature values. Flat Plate Heat Transfer Coefficient During this experiment, I determined the heat transfer coefficient across a flat plate at four different velocities each by three different methods. The first was to measure it 38 experimentally. The second method involved using the theoretical analysis presented in the theory section of this report to come up with a predicted estimate of the heat transfer coefficient. Estimating the heat transfer coefficient with the final method is a direct result of comparisons of the experimental and predicted results. It involves a minor adjustment to the theoretical prediction equation to account for turbulent mixing. I will begin to explain my results for the flat plate by comparing the first two methods of estimating the heat transfer coefficient, experimental and predicted. In order to determine the experimental heat transfer coefficient, I measured the power output, the area of the plate, the steady state temperature of the air, and calculated the steady state temperature of the plate, during the lab. From this data, I used Equation 5 to determine the average heat transfer coefficient over the plate for all four experiments. The second calculation involved using a theoretical analysis assuming laminar flow of the air over the flat plate. During the lab trials, I recorded the velocity of the air, and calculated the film temperature of the air. With this data, I was able to use Equation 60 to calculate the average heat transfer coefficient over the plate for the three forced convection runs and using Equation 77 for the natural convection prediction. Table 2 shows the results of these calculations. Table 2: Heat transfer coefficients along flat plate (W/sqm deg C) Flat Plate: Air Velocity(m/sec) Experimental Experimental Error Theory Percent Error in Predictions 0.0 26.6 12.1% 6.8 292.8% 2.5 39.4 7.7% 22.0 79.6% 5.0 53.6 6.4% 30.9 73.7% 10.0 78.1 1.4% 44.4 75.9% Table 2: Heat transfer coefficients along flat plate. If one were to examine the table , they would be able to note that both experimental and predicted results follow the same trend of increasing with increasing velocity. In order to more clearly see the effect that velocity has on the heat transfer coefficient, it is convenient to graph the data in Table 2. This is shown in Figure 13 below. 39 Heat transfer coefficients versus velocity for laminar flow over flat plate 90 h (Watts per meter squared degrees Celsius) 80 70 60 50 Predicted Experimental Linear (Predicted) 40 30 20 10 0 0 0.5 1 1.5 2 2.5 3 3.5 (Velocity (meters per second))^.5 Figure 13: Heat transfer coefficients versus velocity for flat plate. From this graph, it is clear to see that both the experimental and predicted heat transfer coefficients increase almost linearly with the root of air velocity. Also, from Figure 13, it is obvious that the theoretical predicted equation is not providing good results for our experimental conditions. This conclusion led to the third method of prediction, turbulent mix predictions. This prediction method is exactly the same as the original theoretical prediction method except that I use a different coefficient in Equation 60. The new predicting equation for this method is shown below, (86) Nu hL 1.169 Pr1/ 3 Re x1/ 2 . k Using this new method for my predicted heat transfer coefficients and comparing it to my experimental measurements are shown below in Table 3. 40 Table 3: Heat transfer coefficients along flat plate (W/sqm deg C) Flat Plate: Air Velocity(m/sec) Experimental Experimental Error Theory Percent Error in Predictions 0.0 26.6 12.1% 0.0 #DIV/0! 2.5 39.4 7.7% 38.7 2.0% 5.0 53.6 6.4% 54.4 -1.4% 10.0 78.1 1.4% 78.2 -0.1% Table 3: Heat transfer coefficients along flat plate. The most important thing to note from this table is the large decrease in percent error in the predictions. With this new equation, we have obtained an average error of only 1%. In order to find the functional relationship of the heat transfer coefficient and the air velocity, it is more suitable to see the data from Table 3 in a graph. This is shown in Figure 14 below. Heat transfer coefficients versus velocity for laminar flow over flat plate 90 80 h (Watts per meter squared degrees Celsius) 70 60 50 Turbulent Mix Predicted Experimental Linear (Turbulent Mix Predicted) 40 30 20 10 0 0 0.5 1 1.5 2 2.5 3 3.5 -10 (Velocity (meters per second))^.5 Figure 14: Heat transfer coefficients versus velocity for flat plate. Figure 14 above, illustrates that the new predictive equation also follows the trend of heat transfer coefficients increasing linearly with the root of air velocity. Also from this graph, 41 one is able to visualize that our experimental data lines up well with the new predicted data. Finned Plate Heat Transfer Coefficient The analysis of the finned plate is similar to that of the flat plate. Once again, we will compare three determinations of the heat transfer coefficient. The first will be via experiment. We will measure the temperature of the air, the steady state temperature of the plate, the area of the plate, and the power input to the plate. This data however is not enough to solve for the heat transfer coefficient for the fins. According to Equation 83, we also need the heat transfer coefficient of the plate. I am going to presuppose that this value is equal to the value found for the flat plate at the same velocity in the previous set of experiments. Thus with all of this data, we are able to solve for our experimental value. Next, we will calculate the value of the heat transfer coefficient as predicted from theory for laminar flow over the plate. This calculation is exactly the same as it was for the flat plate, with the use of Equation 60. The data for this part of the experiment is shown below in Table 4 Table 4: Heat transfer coefficients along finned plate (W/sqm deg C) Finned Plate: Air Velocity(m/sec) Experimental Experimental Error Theory Percent Error in Predictions 0.0 5.1 20% 6.6 -22.9% 2.5 32.1 13.2% 22.0 45.7% 5.0 41.5 11.8% 31.1 33.2% 10.0 55.8 2.1% 44.0 26.6% Table 4: Heat transfer coefficients along finned plate. It is important once again to note the trend and the percent error from this table. As before both the experimental and predicted heat transfer coefficients increase with increasing air velocity. In addition, a large error is seen in our predictions from theory and our experimental results. In order to determine the effect of air velocity on the heat transfer coefficient, it will be more convenient to see the data in Table 4 in graphical form. This is presented below in Figure 15. 42 Heat Transfer Coefficients for flow over Finned Plate h (watts per meter squared degree celcius) 70 60 50 40 Predicted Experimental 30 Linear (Predicted) 20 10 0 0 0.5 1 1.5 2 2.5 3 3.5 (Velocity (m eters per second))^.5 Figure 15: Heat transfer coefficients versus velocity for finned plate. Two things become apparent from Figure 15. First, both the experimental data and predictive data increase with increase velocity and second, our predictions from Equation 60 once again do not provide satisfactory results. Consequently, I tried to change the coefficient from Equation 60 and create a new turbulent mix predicting equation shown below. (87) Nu hL .90 Pr1/ 3 Re x1/ 2 k Note the newly developed predicting correlations, Equations 86 and 87 are not the same. The next step was to compare the results obtained experimentally with those obtained from Equation 87. The results of these calculations are shown below in Table 5. 43 Table 5: Heat transfer coefficients along finned plate (W/sqm deg C) Finned Plate: Air Velocity(m/sec) Experimental Experimental Error Theory Percent Error in Predictions 0.0 5.1 20% 0.0 #DIV/0! 2.5 32.1 13.2% 29.9 7.4% 5.0 41.5 11.8% 42.2 -1.8% 10.0 55.8 2.1% 59.7 -6.7% Table 5: Heat transfer coefficients along finned plate. From this table, it should be noted that our error in predictions have dropped substantially, as we now only have an average error of 6%. We would like to again graph this data such that we can more easily see the relationship between the heat transfer coefficient and the air velocity. This is shown below in Figure 16. Heat Transfer Coefficients for flow over Finned Plate 70 h (watts per meter squared degree celcius) 60 50 40 Turbulent Mix Predicted Experimental Linear (Turbulent Mix Predicted) 30 20 10 0 0 0.5 1 1.5 2 2.5 3 3.5 (Velocity (meters per second))^.5 Figure 16: Heat transfer coefficients versus velocity for finned plate. From this figure, it should be noted that our newly predicted heat transfer coefficients follow the same trend of increasing linearly with the root of air velocity. And again, our experimental data line up quite well with our newly predicted data. Effectiveness of Fins 44 To satisfy my last objective, I need to compare how much the addition of fins helped with the heat transfer from a plate. The easiest way to accomplish this is to run the same amount of power into the flat plate and the finned plate, then compare the difference between the temperature of the surface with the temperature of the inlet air for each. In order to analyze this data, it should be noted that at a constant q, h flat Aflat T flat h fin Afin T fin . Thus a smaller T would indicate that better heat transfer is being established because more of the heat being pumped in must be being carried away. Thus in Figure 17 below, I have graphed my T ’s for each run from the fin and the plate. Overal Heat Transfet Com parison (a) 80 70 Delta T (degrees Celsius) 60 34 deg 50 Flat Plate 40 Finned Plate 30 40 deg 20 29 deg 18 deg 10 0 0 2 4 6 8 10 12 Velocity (m eters per second) Figure 17: Temperature Differences recorded for flat and finned plates. Figure 17 reveals that in all cases the fin has increased the heat transfer rate. However, another important aspect of Figure 17 is the decreasing difference of T fin with T flat as 45 velocity increases. This implies that the fin is becoming less effective at increasing the heat transfer rate at higher velocities. Discussion of Results: Effectiveness of Air Velocity on Heat Transfer Coefficient The first objective of my experiment was to find the effect that increasing air velocity would have on the heat transfer coefficient, both for a flat plate and a finned plate. Figures 13 and 15 reveal the answer to this correlation. The heat transfer coefficient increases linearly with the square root of air velocity. This relationship agrees with the theoretical correlation produced by Equation 60. This result is logical. An increase in the free stream velocity, will increase the velocity of the air in the boundary layer. Let us imagine we are an air molecule near the surface of the plate in the boundary layer. At the beginning of the boundary layer, we are floating by at a temperature close to that of the free stream temperature. As we float by at a lower temperature than the plate, we cause heat transfer from the plate to us. This will result in our temperature increasing, which means that we as we move down the plate, our temperature gradient becomes less and we are causing less transfer. Thus at 10cm down the plate, we are causing significantly less heat transfer. However, if we increase our velocity, then we will be effecting the same temperature gradient in the beginning but not sit at the same point long enough to absorb as much heat. Instead, we will be replaced more quickly by the air molecules behind us at the same temperature. Thus, our temperature gradient 10 cm down the plate will not be as diminished as it was at the lower velocity. The net result of this will increase the heat transfer coefficient. This result was in agreement with both the experimental and the theoretical predictions. The significance of these results is most evident in the design of heat exchangers. If we are operating a heat exchanger to heat up fluid A by using fluid B, but it is noted that we need to increase the outlet temperature of fluid A by more than its current value, then we 46 must increase the amount of heat transferred. From the results above, we know that if we increase the velocity of fluid B, then the average heat transfer coefficient will also rise, allowing us to transfer more heat to fluid A and thus raising its outlet temperature. Comparison of Experimental results to theoretical Forced Convection The second objective of my experiment is to compare my results obtained from these experiments with those predicted by theory. From Tables 2 and 4, we see that the predicted results are not in agreement with the experimental results. For the flat plate, ignoring natural convection, our average prediction error was around 80%, while for the finned plate our average error did drop, but still remained relatively high at a value of 35%. Comparing this to the average experimental error which stayed relatively low for both setups at an average of 6%, it would appear that the predictive equations do not give excellent results. However, this may not necessarily be the case. Looking back at Tables 2 and 4, or by looking at Figures 13 and 15, we can see a couple of important characteristics. First, the error in the prediction seems to diminish as we increase the velocity. I rationalized this information to mean that as the flow becomes more turbulent from a Reynolds number standpoint, though still laminar, then the predicting equation produced better results. Next, we see a large decrease in experimental error when we move from the flat plate to the finned plate, however, the error is still substantial. I interpreted this to be the fact that the boundary layer was forming over a plate that was separated from the surface of the chimney, extending away from it. As a result of combining the above observations, I came up with the following conclusion. Because the flat plate is along the level of the chimney, though a bit raised, then we must be starting a boundary layer along the chimney before even reaching the flat plate. Of course ,only the flat plate is heated thus we would have to use Equation 60b to find the average heat transfer coefficient. However, without doing the math, we can see that this cannot fully explain what is happening. If this were truly the case ,then from Equation 42, we know the boundary layer will be thicker over the plate then it would 47 have been from our original assumption of the boundary layer starting at the leading edge of the plate. But a thicker boundary layer would push the free stream temperature farther away from the plate, making it harder to transfer heat. Therefore this result would only lower the predicted heat transfer coefficient and thus increase our error. Something else must be happening. It is my hypothesis that the boundary layer is starting along the chimney, however, because the flat plate is a bit raised, it is causing turbulence in the boundary layer when it reaches the flat plate. This may cause the mixing in the boundary layer to be turbulent even though it is still at a low Reynolds number as shown below in Figure 18. Turbulence begins because the boundary layer formed over the duct wall hits the edge of the flat plate which is slightly raised, thus disturbing the molecules in the boundary layer to form a turbulent mix u8 u8 T8 A8 u8 Duct Wall Figure 18: Turbulence over flat plate. This is not a far stretch to explain my results.. As people have run the Reynold’s experiment described earlier and found turbulence at low Reynolds numbers by adding factors to increase mixing. Thus if we experience better mixing in the boundary layer, we will increase the heat transfer coefficient because we will ensure that molecules farther up in the boundary layer at a lower temperature are being forced closer to the plate. This would increase the average temperature gradient close to the plate and result in higher experimental heat transfer coefficient as compared to if the flow was perfectly laminar. 48 To test this theory with my data I had to make a conjecture. Because the Reynolds probably remains at the laminar level, assuming the velocity of molecules in the turbulent mixing boundary layer haven’t changed much, than the correlation of Equation 60 that the Nusselt number is a function of Reynolds to the one half power and Prandtl to the one third power probably remains true. However, due to the turbulent mixing I feel that the coefficient of this equation should increase, because the increase of the average temperature gradient near the plate should proportionally raise the heat transfer coefficient. I fitted the data using (88) Nu hL function(Pr1/ 3 Re L1/ 2 ) Constant Pr1/ 3 Re x1/ 2 . k From the analysis of the flat plate I obtained, (86) Nu hL 1.169 Pr1/ 3 Re x1/ 2 . k I then extended this analysis to the finned plate. For this situation, the boundary layer could not have started early because the fins are protruding into the air with no surface before them. However, as a result of their blunt, non aerodynamic shape, they might be causing the incoming air to also form a turbulent mix boundary layer. When the air hits the fins, as shown in Figure 19 below, it will cause the stream lines to form eddy circles which result in a turbulent mix at the low Reynolds number. 49 Turbulence begins because the incoming air will h it the blunt side of the fin, causing the mo lecules to be disturbed in many different directions. This causes the boundary layer to have a slight turbulent mix. u8 T8 A8 u8 Figure 19: Turbulence around fins. Thus, I performed the same analysis as before using Equation (88) for the finned plates I obtained, (87) Nu hL .90 Pr1/ 3 Re x1/ 2 k Returning to the comparison of the experimental results with the new predictive equations using turbulent mixing, we have Figures 14 and 16 as well as Tables 3 and 5. Our new predictive equations yield us excellent results when compared with those obtained from experiment. With so few data points, it is hard for me to tell whether these equations are truly accurate. However, a couple of features of these equations do give me confidence in declaring these equations as a more accurate model to what is happening in this lab. First, both of the constant increased. This agrees with my theory that mixing is turbulent and is increasing heat transfer proportionately. Second, the constant for the flat plate is larger than the constant for the small plate. This is logical because they are both experiencing a different type of situation when they become a turbulent mix. The flat plate has a built up 50 boundary layer which is being distorted allowing it to already be built up on the plate as a turbulent mix layer. However, the fins boundary layer only forms once it hits the fins’ leading edge, causing it to alleviate some of the turbulence into the free stream air above the plate instead of retaining it all in the boundary layer. Subsequently, the fins boundary layer will be less of a turbulent mix and thus it has a lower coefficient. In summary, the theoretically derived predictive equations, do not provide excellent results. However, the defining equations used to develop these equations, Equation 88, which led to Equations 86 and 87, does provide excellent results. Therefore, although we probably cannot increase accuracy above 30% without building the exact situation, we can quickly develop equations which can achieve an accuracy of about 15% by running a simple convection experiment. Natural Convection Again, looking back at Tables 2 and 4, or by looking at Figures 13 and 15, we see that our predictive equations for natural convections, also, do not provide satisfactory results. The error over the flat plate is about 300%, while the error for the finned plate was close to an acceptable value of -22%. In all probability, these percent errors are not giving the true accuracy of the equation because of only one data point for each measurement. But from the large error over the flat plate, it is my conclusion that the predictive equations for natural convection will not provide good results. I don’t have an answer as to why the predictive results do not match the experimental results. My only guess would be the assumptions used to create the natural convection models probably aren’t applicable to all situations. Therefore if an engineer wishes to use natural convection in his process, he should try to determine the heat transfer coefficients from small scale experiments of his situations rather than use the predictive equations. However, these conclusions are only a result of one data point for each situation and thus I would need more data to make a more solid judgment. On the other hand, this doesn’t mean that the results from this experiment are useless. Although, we weren’t able to accurately predict the values of the heat transfer coefficient at natural convection, we were able to prove that natural convection is present. Our 51 experiment has concluded that the limiting value to convection is not zero at zero velocity as forced convection theory predicts. but that heat transfer still takes place, and at a sufficient level, when there is a zero velocity atmosphere. The significance of the above results is to enlighten engineers about natural convection. If they are armed with this information they would be prepared and can account for its presence in their design if it calls for a piece of equipment at a certain temperature. Or more importantly, if their design only requires a small amount of heat transfer to keep an electronic device at a low temperature, then they can save money by not installing a pump to create forced convection, but instead let the density gradient due the work for free in creating the necessary heat transfer rate. Effectiveness of Adding Fins to a Flat Plate My final objective was to determine how much the fins helped in the heat transfer from the flat plate at the same velocity. From Figure 17, we see the temperature differences for both the finned plate and the flat plate. Since the heat being convected away from each is a constant, the lower T , signifies a greater heat transfer rate has been achieved. From the first two data points, it appears that the fins become more effective moving from natural convection to forced convection. With the data from this figure, we see that for every run the fins provide a faster heat transfer rate than the flat plate. However, this is fairly obvious because we are adding more area for heat transfer. Another, and more important issue, is the trend we see in the effectiveness of the fins. For all of our runs at constant q, we have, h flat Aflat T flat h fin Afin T fin . For the fins, we have increased area to increase the heat transfer rate. This provided us with a lower T . However, as our air velocity increased, the increase in both h flat and h fin , to obtain higher heat transfer rates overshadowed the increased area provided by the fins. Thus, the effect of the extra area provided by the fins dropped. Accordingly as the air velocity increased, the heat transfer rate of the flat plate greatly increases, however the heat transfer rate of the fins does increase, but not as dramatically. Therefore, the effectiveness of adding the fins drops as we increase the air velocity. 52 If an engineer is trying to decide whether he should add the extra cost of installing fins to his surface, the variable he should look at is his operating fluid velocity. If the velocity is high, the gain in heat transfer rate will probably not be worth the cost. However, if he is operating at a fairly low fluid velocity, it will probably be worth the cost to add the fins, because the increase in heat transfer will be substantial. Conclusions: This experiment provided four results that are very important in the life of a process or a design engineer. First, this experiment proved that the heat transfer coefficient increased linearly with an increase in the square root of air velocity. This is shown in Figures 13 and 15. The significance of this result is most obvious in the design of heat exchangers. If the engineer has an old exchanger that he would like to use and performs all of the calculations at a given flow rate but finds that he is not providing enough fluid to heat up the cold product, he now knows that he only needs to increase the flow rate of the heating fluid to increase the heat transfer between them. However, I should note that an increase in velocity usually also means an increase in pressure drop, the engineer should be aware of this in making his design. Secondly, this experiment has proven the theoretically derived equations for convection but have large errors associated with them in trying to predict the absolute values of the heat transfer coefficient if not applied to the exact geometry described with their assumptions as shown in Tables 2 and 4. Nonetheless, the equations are useful for predicting the trend. It is recommended to first run small scale experiments to obtain a constant fudge factor which can be applied to the correlation stated in the predictive equations. This was shown to reduce error to an acceptable level as shown in Tables 3 and 5. The significance of this result is it allows engineers to use correlations of heat transfer coefficients for the design of heat exchangers. It would be too expensive for the engineer to try multiple heat exchangers, run tests and pick the right one. Thus, the engineer must run through the equations derived for convection heat transfer in heat exchangers to predict the type, length, and area of his heat exchanger. This data proves that such correlations will yield acceptable results. 53 The third major result of this experiment is the effectiveness of fins. At all velocities, this extended surface provided a great heat transfer rate. Therefore, if the engineer wishes to achieve a large heat transfer rate, but is limited in his length/area requirement, he will be able to simply add fins to the surface. However, he should note that the effectiveness of the added fins decreases as the velocity increases, as is shown in Figure 17. The significance of this result comes into play when an engineer is designing a heat exchanger which must dissipate a certain amount of heat to the surroundings. Before he chooses his design, the engineer should check the velocity of the flowing fluid. If the velocity is high, the added fins will probably not be worth the cost. If however, the velocity is low, the cost of adding the fins will be overshadowed by the large increase in the heat transfer rate. Finally, the last result is that the heat transfer coefficient is a function of geometry only not Area. This was first noted in the theory section when Area dropped out and all of the correlations were made for similar geometries. This was, again, proven in the experiment when the same predictive equation, Equation 60 was used for the flat plate, at an area of .011 m 2 and for the fins of the finned plate, each at an area of .0068 m 2 . Both of the adjusted equations based from Equation 60 provided great results, thus the geometries of the flat plate and the fins (flat surface) allowed us to use the same equation, the difference in their areas had no effect. 54 Recommendations: In my experiment I had trouble gaining too much confidence in my conclusions because of the lack of data points that I was able to attain. Thus for the next time this experiment is run I have a few suggestions for the experimenter. First he might want to concentrate on getting all 8 runs on the flat plate to attain enough data to ensure himself of the correlations. Or he might want to keep the same 4 runs on a flat plate, 4 runs on an extended surface, but instead try to make both the fins and the flat plate, more aerodynamic thus to ensure a fully laminar boundary layer will develop. 55 Nomenclature: A=Area=(m 2 ) J Cp heat capacity= 0 kg C f =Fin Gr=Grassof Number m g=Acceleration due to gravity= 2 s W h=heat transfer coefficient= 2 0 m C W k=Conductive heat transfer coefficient= 0 m C L=Length=(m) Nu=Nusselt number N P=pressure= 2 m Pr=Prandtl number q=rate of heat transfer=(W) Ra=Rayleigh number Re=Reynolds number s =surface T=Temperature= 0 C t=time=(s) t =thermal m u=velocity= s V=Volume= m 3 w =wall x=vertical position=(m) y=horizontal position=(m) 56 m2 α=thermal diffusivity= s β=volumetric thermal expansion coefficient= K -1 δ=boundary layer thickness=(m) η=fin efficiency kg μ=viscosity= ms m2 ν=momentum diffusivity= sec θ=T-T = 0 C kg ρ=density= 3 m N 2 m shear stress= m3 υ=specific volume= kg Δ=Change in a variable=final-initial =Property at free stream, far removed from system Average Property 57 References: 1. Centre for the Analysis and Dissemination of Demonstrated Energy Technologies. Energy Conservation in the Pulp and Paper Industry. CADDET Analyses Series. March 28, 2001. 2. Holman J.P. Heat Transfer 9th Edition. McGraw Hill, Inc. (2002) 3. Icoz and Jaluria. Rutger University. Design of Cooling Systems for Electronic Equipment Using Both Experimental and Numerical Inputs. Journal of Electronics Packaging. May 31, 2004. 4. McCabe, W. L. and Smith, J. C. Unit Operations of Chemical Engineering, 3rd Edition, McGraw-Hill Book Company, 1976. 5. Principles of Flow. Ed. Alicat ScientificTM. 5 July 2005. http://www.alicatscientific.com/flow_principles.php 6. Welty, J.R., et al. Fundamentals of Momentum, Heat, and Mass Transfer – 4th Ed. John Wiley & Sons, Inc. New York. (2001) 58 APPENDIX: Appendix 1: Sample Calculations Run1: Flat Plate Low velocity Given Variables: Area .011 m 2 measured Length .1 m measured W tabulated m 0C q 20 W measured k plate 166.5 u 2.5 m measured s Experimental Data Data Points Time (min) Air InletTemperature (°C) Plate Surface Temperature (°C) exp(-kt) K 1 0 25.7 1.00 0.05 2 5 25.7 35.5 0.76 R Squared 3 10 25.7 45.7 0.58 4 15 25.7 51.2 0.44 5 20 25.7 56 0.33 6 25 25.7 59.5 0.25 7 30 25.7 62.3 0.19 8 35 25.7 64.6 0.15 9 40 25.7 66.3 0.11 10 45 25.7 67.8 0.08 11 50 25.7 69.1 0.06 12 55 25.7 70 0.05 1.00 Table 6: Experimental data obtained for flat plate, at 2.5 meters per second. Calculating Steady State Temperature: The time, air inlet temperature, and the plate surface temperature were measured during the lab. Then a K value was assumed to be one. Then data was calculated of the form exp K * t . Then the Temperature of the Surface was graphed versus exp K * t and then sent through a linear regression through excel. Then 59 K was goalseeked until the correlation coefficient was closest to unity. The graph below shows the final result of this calculation. Temperature of Surface (deg C) Establishing Steady State 80 70 60 50 40 30 20 10 0 0.00 Series1 Linear (Series1) y = -47.016x + 71.8 R2 = 0.9981 0.20 0.40 0.60 0.80 exp(-K*t) Figure 20: Steady state calculation for the first run. From Equation85, we calculate, Ts , steady state 71.80 C Calculating Film Temperature T film Ts , steady state T 2 71.8 25.7 48.75 2 Calculating Properties at Film Temperature: W tabulated m0C W kair .031 0 tabulated mC kg air 1.94 105 tabulated ms kg air 1.102 3 tabulated m Pr f .703 tabulated k plate 166.5 Calculating Experimental h: From Equation 1 we have, 60 q h(Ts T ) A q h A(Ts T ) h 20W .011m 2 71.8 25.7 C 0 hexp 39.44 W m 0C Calculating Predictive h: From Equation 8 we have Re L uL kg m 2.5 .1m 3 m s 14394 5 kg 1.94 10 ms 1.102 And from Equation 60 hL .664 Pr1/ 3 Re L1/ 2 k .664 Pr1/ 3 Re L1/ 2 k h L Nu .664 .703 1/ 3 14394 h hlaminar 21.96 for Re 5 1010 x0 0 1/ 2 W .031 0 m C .1m W m 2 0C Calculating Predictive Error: hexp hlaminar 100% h laminar 39.44 21.96 Prediction Error 100% 21.96 Prediction Error 79.6% Prediction Error Calculating Experimental Error: The error from q is derived from the fluctuations of the value read during the experiment, the error from A was determined based on the accuracy of reading a ruler, and finally the error in temperatures was given from the manufacturers of the thermisters. 61 q hA Ts T h q A Ts T h q A Ts T q A Ts T 1.4 .0005 .1 .1 20 .011 71.8 25.7 h .1207 12.1% h Calculating the turbulent mix coefficient for Nusselt number: Based on the larger error associated with using Equation 60 as my prediction equation, I moved to calculate my own equation based on turbulent mixing over the flat plate. This equation uses the same correlation as equation 60 but has an adjusted coefficient to account for the turbulent mixing. In order to calculate this coefficient, I used the percent error from all three forced convections runs and summed them. Then I sought to minimize this error by guessing different values of the coefficient, using goalseek in Excel. The data from this calculation appears below. Forced Convection Run # New coefficient Re Pr h est h exp error 1 1.168197768 14393.94 0.703 38.63252 39.4399527 0.807435 2 1.168197768 29362.59 0.705 54.33877 53.633682 0.705084 3 1.168197768 60632.94 0.706 78.12178 78.1217878 1.14E-05 1.512531 Table 7: Experimental data obtained for flat plate. Run5: Finned Plate Low velocity Given Variables: 62 Area plate .011 m 2 measured Length .1 m measured Length fins .1m (measured ) Depth fins .068m (measured ) Perimeterfins .206m (measured ) W tabulated m 0C q 21 W measured k plate 166.5 m measured s W 39.44 2 0 calculated m C u 2.5 hplate Calculating heat transfer coefficient and efficiency From our previous development we have, Equations 82 and 83, q h0 A0 h f A f f f m2 T T s hPkA 0 tanh(mL) tanh(mL) hPL 0 mL hf P kA where A0 is the free are of the flat plate. This area is the total area of the flat plate minus the area of the 9 fins sitting on the flat plate. Thus we have, A0 Aplate Af .011 9*.0004 .0074 The only unknown in the above equations is hf, however, these equations are too complex to try and solve for hf explicitly. The calculation of hf is therefore a trial and error procedure solving all three of these equations simultaneously. Thus we have, 63 W 0 21W 39.44 2 0 .0074m 2 h f .1224m 2 f 31.06 25 C m C h f .206m .1m) W 2 166.5 0 .1224m mC h f .206m .1m W 2 166.5 0 .1224m mC tanh( f And from Excel we obtain, h f 32.08 W m 2 0C f .836 Fin Effectiveness for Runs 1 and 5 Given Variables: Ts , fin 31.10 C T , fin 250 C Ts , flat 71.80 C T , flat 25.10 C Effectiveness of adding fins: T fin Ts , fin T , fin T fin 31.10 C 250 C 6.10 C T flat Ts , flat T , flat T flat 71.80 C 25.10 C 46.10 C Effectiveness compares T flat T fin 46.10 C 6.10 C 400 C Reviewed and Validated by: Thomas Salerno ____________________________ Jennifer DiRocco ____________________________ Greg Rothsching ____________________________ Stephen Johnson ____________________________ 64 Appendix 2: Raw Data Run 1: Flat Plate Low Velocity Air Speed (m/sec) Power Input (Watts) Area (m*m) 2.5k plate 20L plate (m) 0.011Air Viscosity Steady State Values 166.5Air Density 0.1Pr 0.000019423Tf 1.102 0.703 48.75 Plate Surface Temperature Time (min) Air InletTemperature (°C) (°C) Fin T1 (°C) Fin T2 (°C) 25.7 71.8 +-0.5 Data Points Experimental Data Data Points Plate Surface Temperature Time (min) Air InletTemperature (°C) (°C) 1 0 25.7 exp(time) K 1.00 0.05 2 3 5 10 25.7 25.7 35.5 45.7 0.76 R Squared 0.58 1.00 4 5 6 7 8 9 10 11 12 15 20 25 30 35 40 45 50 55 25.7 25.7 25.7 25.7 25.7 25.7 25.7 25.7 25.7 51.2 56 59.5 62.3 64.6 66.3 67.8 69.1 70 0.44 0.33 0.25 0.19 0.15 0.11 0.08 0.06 0.05 Temperature of Surface (deg C) Establishing Steady State 80 70 60 50 40 30 20 10 0 0.00 Series1 Linear (Series1) y = -47.016x + 71.8 R2 = 0.9981 0.20 0.40 0.60 0.80 exp(-K*t) 65 Run 2: Flat Plate Med Velocity Air Speed (m/sec) Power Input (Watts) Area (m*m) Data Points 5k plate 20L plate (m) 0.011Air Viscosity Steady State Values 166.5Air Density 0.1Pr 0.00001914Tf 1.124 0.705 41.85 Plate Surface Time (min) Air InletTemperature (°C) Temperature (°C) Fin T1 (°C) Fin T2 (°C) 24.9 58.8 Experimental Data Data Points Plate Surface Time (min) Air InletTemperature (°C) Temperature (°C) exp(time) 1 0 25 69.6 1 0.0581 6 5 24.9 66.3 0.747888 11 10 63.9 0.559337 0.995365 16 15 63.1 0.418322 21 20 25.1 61.9 0.312858 26 25 61.2 0.233983 31 30 60.6 0.174993 36 35 60.1 0.130875 41 40 59.7 0.09788 46 45 24.8 59.5 0.073203 51 50 59.3 0.054748 56 55 59.2 0.040945 67 66 65 y = 9.821x + 58.805 R2 = 0.9954 64 63 Series1 Linear (Series1) 62 61 60 59 58 0 0.2 0.4 0.6 0.8 66 Run 3: Flat Plate High Velocity Air Speed (m/sec) Power Input (Watts) Area (m*m) Data Points Data Points 10k plate 166.5Air Density 1.14475 19L plate (m) 0.1Pr 0.706 0.011Air Viscosity 0.00001888Tf 36.055 Steady State Values Air InletTemperature Plate Surface Time (min) (°C) Temperature (°C) Fin T1 (°C) Fin T2 (°C) 25 47.11 Time (min) 1 6 11 16 21 26 31 36 41 46 51 56 61 Experimental Data Air InletTemperature Plate Surface (°C) Temperature (°C) 0 24.9 5 10 15 25.1 20 25 30 25.2 35 40 45 50 55 60 exp(time) 58.5 53.1 50.7 49.2 48.2 47.8 47.5 47.3 47.2 1 0.110469 0.575598 0.331313 0.999026 0.190703 0.109768 0.063182 0.036368 0.020933 0.012049 0.006935 0.003992 0.002298 54 53 y = 10.53x + 47.111 R2 = 0.999 52 51 Series1 50 Linear (Series1) 49 48 47 46 0 0.2 0.4 0.6 0.8 Run 4: Flat Plate Natural Convection Air Speed (m/sec) Power Input (Watts) 0k plate 20L plate (m) 166.5Air Density 0.1Pr 1.06025 0.701 67 Area (m*m) 0.011Air Viscosity Steady State Values 0.00001996Tf Plate Surface Time (min)Air InletTemperature (°C)Temperature (°C) Fin T1 (°C) 25.5 93.8 Data Points 59.65 Fin T2 (°C) Experimental Data Plate Surface Time (min)Air InletTemperature (°C)Temperature (°C) exp(time) 1 0 25.2 47.3 1 0.035255 6 5 55.6 0.838388494 11 10 60.6 0.702895267 0.999246 16 15 21 20 26 25 31 30 36 35 41 40 46 45 51 50 56 55 61 60 Data Points 87 25.5 91.75 0.046553076 37.1 Run 1: Fin Plate Low Velocity Air Speed (m/sec) Power Input (Watts) Area Data Points Data Points 2.5Efficiency Estimated Te 0.236445618Po fin 0.208Ao fin 20Efficiency Calculate 0.836044514Pl fin 0.204Al fin 0.011Area fin 0.1224k fin 166.5L fin Steady State Values Air InletTemperature Plate Surface Fin T2 Time (min)(°C) Temperature (°C) Fin T1 (°C) (°C) Fin T3 (°C) 25 31.062 26.6 26.5 26.2 Experimental Data Air InletTemperature Plate Surface Fin T2 Time (min)(°C) Temperature (°C) Fin T1 (°C) (°C) Fin T3 (°C) 1 0 6 5 25.5 28.5 11 10 30.1 16 15 30.7 21 20 30.9 26 25 31 31 30 26 31 26.6 26.5 26.2 36 35 41 40 46 45 51 50 68 56 61 55 60 31.5 31 30.5 30 Series1 y = -6.4873x + 31.062 R2 = 0.9992 29.5 Linear (Series1) 29 28.5 28 0 0.1 0.2 0.3 0.4 0.5 Run 2: Fin Plate Med Velocity Air Speed (m/sec) Power Input (Watts) Area Data Points Efficiency 5Estimated 0.350340136Po fin 0.208Ao fin 20Efficiency Calculate 0.799479113Pl fin 0.204Al fin 0.011Area fin 0.1224k fin 166.5L fin Steady State Values Air Time InletTemperature Plate Surface Temperature Fin T1 Fin T2 Fin T3 (min) (°C) (°C) (°C) (°C) (°C) 25 29.9 26.8 26.75 26.6 Experimental Data Air Time InletTemperature Plate Surface Temperature Fin T1 Fin T2 Fin T3 Data Points (min) (°C) (°C) (°C) (°C) (°C) 1 0 31 6 5 29.9 11 10 29.9 16 15 26.4 29.9 21 20 29.9 26.8 26.8 26.6 69 26 31 36 41 46 51 56 61 25 30 35 40 45 50 55 60 Run 3: Fin Plate High Velocity Air Speed (m/sec) Power Input (Watts) Area Data Points Data Points 10Efficiency Estimated 0.418492739Po fin 0.208Ao fin 20Efficiency Calculate 0.750989434Pl fin 0.204Al fin 0.011Area fin 0.1224k fin 166.5L fin Steady State Values Air Time InletTemperature Plate Surface Fin T1 Fin T2 Fin T3 (min) (°C) Temperature (°C) (°C) (°C) (°C) 25 28.879 26.7 26.6 26.57 Time (min) 1 0 6 5 11 10 16 15 Experimental Data Air InletTemperature Plate Surface (°C) Temperature (°C) 26.4 Fin T1 (°C) 27.2 28.6 28.8 28.9 26.7 Fin T2 (°C) 26.6 Fin T3 (°C) 26.57 70 29 28.8 28.6 28.4 y = -1.6795x + 28.879 R2 = 0.999 28.2 Series1 Linear (Series1) 28 27.8 27.6 27.4 27.2 27 0 0.2 0.4 0.6 0.8 1 1.2 Run 4: Fin Plate Natural Convection Air Speed (m/sec) Power Input (Watts) Area Efficiency Estimated 0.925287356Po fin 0.208Ao fin Efficiency Calculate 0.969085768Pl fin 0.204Al fin 0.011Area fin 0.1224k fin 166.5L fin Steady State Values Air InletTemperature Plate Surface Data Points Time (min)(°C) Temperature (°C) Fin T1 (°C) Fin T2 (°C) Fin T3 (°C) 25 59.8 58.5 58 55.1 Experimental Data Air InletTemperature Plate Surface Data Points Time (min)(°C) Temperature (°C) Fin T1 (°C) Fin T2 (°C) Fin T3 (°C) 1 0 29.4 6 5 34.5 11 10 38.6 16 15 21 20 26 25 31 30 36 35 41 40 46 45 51 50 71 56 61 55 60 26.4 5 8 . 59.85 58 55.1 72