You are on page 1of 12

Metin Renksizbulut1

e-mail: metin@uwaterloo.ca

Laminar Flow and Heat Transfer in the Entrance Region of Trapezoidal Channels With Constant Wall Temperature
Simultaneously developing three-dimensional laminar ow and heat transfer in the entrance region of trapezoidal channels have been investigated using numerical methods in the Reynolds number range from 10 to 1000. The principal and secondary velocity elds, the temperature eld, and all associated heat and momentum exchange parameters have been examined. The present results for the fully developed ow region of the channels compare well with the available literature. In the entrance region, it is observed that the axial velocity proles develop overshoots near the walls and particularly at the channel corners. It is shown that boundary-layer type of approximations, which lead to Reynoldsnumber-independent Poiseuille and Nusselt numbers, can be used for Reynolds numbers over 50 and after a few hydraulic diameters from the channel inlet. It is also shown that hydrodynamic entrance lengths calculated with methods based on fully developed ow data are grossly in error. New correlations are proposed for the entrance length, and for the friction and heat transfer coefcients. DOI: 10.1115/1.2130405 Keywords: trapezoidal channel, entrance region, friction factor, heat transfer coefcient

Hamid Niazmand
Mechanical Engineering Department, University of Waterloo, Waterloo, Ontario, Canada N2L 3G1

Introduction

Laminar ow and heat transfer in channels of arbitrary cross section have received considerable attention over the years due to their obvious practical importance. Shah and London 1 and Shah and Bhatti 2 have provided comprehensive reviews of a large number of these studies. In recent years, advances in manufacturing technologies, coupled with a drive toward MEMS, have renewed interest in laminar ows in relatively short channels with irregular cross sections. For short channels, the entrance region where the velocity and temperature proles undergo rapid transformations from essentially uniform inlet proles to fully developed parabolic proles is of paramount importance. In the entrance region, an analytical solution of the problem is not feasible due to the nonlinear inertia terms in the momentum equations. Boundary-layer assumptions provide major simplications and have been used extensively. In one of the earliest studies, Schiller 3 analyzed the hydrodynamic entrance region of a tube by radially patching the boundary layer velocity prole near the wall with an inviscid core. Schlichting 4,5 obtained approximate solutions using perturbation techniques and the idea of patching the Blasius boundary layer solution with a perturbed fully developed velocity prole in the axial direction. Both methods have received subsequent renements; see, for example, Schmidt and Zeldin 6 for details. An alternative approach in the analytical formulation of the problem is the linearization of the inertia terms of the NavierStokes equations. Assuming constant pressure over a given cross section and neglecting axial diffusion, Langhaar 7 developed an analytical solution for the entrance length of a circular duct. Similarly, Han 8 obtained an analytical solution for the threedimensional 3D ow in the entrance region of a rectangular duct. Comparable methodology was employed by Lundgren et al. 9 to predict the pressure drop in the entrance region of channels
1 Corresponding author. Contributed by the Heat Transfer Division of ASME for publication in the JOURNAL OF HEAT TRANSFER. Manuscript received February 22, 2005; nal manuscript received July 14, 2005. Review conducted by Anthony M. Jacobi.

of arbitrary cross section. In recent years, numerical methods have been used extensively in the investigation of the hydrodynamic entrance problem. Yet, in many cases the boundary-layer approximations have been applied to the governing equations to reduce the computational effort e.g., 1013. Momentum and heat transfer processes in the entrance region of noncircular channels are relatively complex and, surprisingly, have not been studied in depth despite their importance to some practical engineering devices. For example, short channels with a variety of trapezoidal cross sections are commonly encountered in modern compact heat exchangers. Mini- and microuidic devices, such as sensors, actuators, pumps, etc., all involve transport phenomena in short channels with relatively complex geometries. In the present work, the focus is on trapezoidal channels Fig. 1. Fully developed laminar ows in trapezoidal ducts were analyzed in considerable detail by Shah 14. A discrete least square method was employed to develop series solutions for the velocity and temperature elds. Friction factors and Nusselt numbers for different side angles and aspect ratios were calculated. More recently, Flockhart and Dhariwal 15 used the same technique along with an experimental investigation to study pressure drop in microscale trapezoidal ducts with hydraulic diameters of 50 120 m. The microchannels were fabricated by an etching technique resulting in trapezoidal cross sections with side angles of = 54.74 deg silicon substrate. It was concluded that a formulation based on the conventional no-slip boundary conditions no velocity slip, no temperature jump adequately predicts the experimentally observed ow characteristics. Sadasivam et al. 16 also considered laminar fully developed ows through singleand double-trapezoidal hexagonal channels using a nitedifference method. Solutions for the velocity and temperature elds were obtained for a wide range of duct aspect ratios and side angles. Recently, Morini 17 numerically analyzed fully developed ows in rectangular and trapezoidal microchannels. However, for the latter, only one side angle = 54.74 deg was considered. The calculated friction coefcients for a wide range of channel aspect ratios are shown to agree well with available data. JANUARY 2006, Vol. 128 / 63

Journal of Heat Transfer

Copyright 2006 by ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

Performing calculations to obtain an estimate of the compromising accuracy associated with such an approximation is of its own value and interest.

Problem Formulation

Fig. 1 Flow geometry and the coordinate system

The coordinate system and basic geometric variables that describe a trapezoidal channel are shown in Fig. 1. The aspect ratio and the side angle are denoted by = h / 2a and , respectively, where h is the channel height and a is the half-length of the top wall. The principal ow is in the z direction, and since the ow eld is symmetric with respect to the x-z plane, solution is required only for half of the cross-sectional area. The laminar constant-property viscous ow under study is governed by the usual continuity, momentum, and energy equations as follows:

V =0
The literature discussed thus far addresses thermally and hydrodynamically fully developed laminar ows. Under these conditions, the governing equations basically reduce to two Poisson equations, where, in many cases, analytical solutions can be found. However, much less information is available for developing ows. Lawal and Mujumdar 18,19. studied the 3D laminar incompressible ow of a power-law uid in ducts. They neglected molecular diffusion in the principal ow direction and decoupled the transverse and longitudinal pressure gradients. A variation of Patankar and Spaldings 20 marching procedure as proposed by Briley 21 for this type of parabolic problems was adopted to predict the heat transfer coefcient and pressure drop in square, trapezoidal, and pentagonal ducts. Asako and Faghri 11 and Asako et al. 12 also employed the numerical scheme of Patankar and Spalding 20 together with a coordinate transformation to map the irregular cross section onto a rectangular domain to study the entrance region of polygonal and rhombic ducts. Friction and heat transfer coefcients were obtained for boundary conditions of uniform heat ux and uniform temperature at the walls. Farhanieh and Sunden 22 numerically studied the problem of simultaneously developing uid ow and heat transfer in the entrance region of trapezoidal ducts. A nite volume formulation was used, and calculations were carried out with a uniform wall temperature for several channel aspect ratios and side angles. The secondary ow development, and the axial variation of the friction and heat transfer coefcients were examined in detail. However, the axial development of the ow eld with its distinct overshoot feature was not addressed. One of the interesting features of the ow development from a uniform inlet velocity prole is the emergence of overshoots in the axial velocity prole such that the maximum local velocity occurs near the walls as opposed to the channel core. This effect is a function of Reynolds number and becomes more pronounced at higher Re. Furthermore, the effects of Reynolds number on the key ow parameters were not studied. It should be noted that the ow development becomes independent of Reynolds number only when boundary-layer type of approximations are adopted axial diffusion and cross-ow pressure gradients ignored. Finally, no information regarding the entrance length was provided. In the present work, the momentum and energy equations for 3D laminar ows in the entrance region of trapezoidal channels are solved using a control-volume-based numerical method in the Reynolds number range from 10 to 1000. Different channel side angles and aspect ratios are considered, and the results are discussed in detail. Practical engineering correlations for the entrance length, friction factor, and heat transfer coefcient are also provided. It is worth emphasizing that, in the present study, molecular diffusion of heat and momentum in the principal ow direction are included in the analysis. As indicated earlier, a parabolic treatment for the governing equations has been common practice to simplify the computational process, and therefore, most of the available literature is restricted to high Reynolds number ows. 64 / Vol. 128, JANUARY 2006

1 2 3

DV = p + 2V Dt c p
DT = k 2T Dt

is the velocity vector, T is temperature, is density, p is where V pressure, is dynamic viscosity, c p is specic heat, and k is thermal conductivity. Inow boundary conditions correspond to a uniform at velocity prole such that u = v = 0, and w = Wi, with a uniform inlet temperature T = Ti. For outow, zero gradients along the axial ow direction are applied to all variables, f / z = 0, where f = u , v , w , T. The channel walls are uniformly at a temperature of Tw, where the ow satises the no-slip boundary condition, u = v = w = 0. Since the problem under study is considered as a massdriven ow, the pressure is set equal to zero at the outlet, while zero gradients are assigned at all other boundaries including the inlet. The Reynolds number Re= WiDh / is based on the uniform inlet velocity and the hydraulic diameter Dh dened as four times the cross-sectional ow area over the wetted perimeter. It is also specied that Pr= k / c p = 1, where Pr is the Prandtl number.

Solution Method

The numerical methodology is based on a transformation of the governing equations into a generalized nonorthogonal coordinate system. The equations are solved using a projectionlike method originated by Chorin 23 and developed further by Dwyer 24 and the authors. The velocity components are rst calculated from the momentum equations using an alternating-direction predictorcorrector scheme based on the pressure eld at the previous time step. Then the pressure correction is calculated from an equation designed to satisfy conservation of mass directly. This pressure correction equation is a Poisson-type equation solved with a matrix-free and preconditioned version of GMRES generalized minimal residual algorithm 25. This method is capable of enhancing the convergence rate of the Poisson solver as compared to the traditional successive-overrelaxation solvers. Typical slow convergence of the pressure elds in duct ows is largely caused by the high aspect ratios of the computational cells that result from the ow geometry. Further enhancement in convergence rate of the solution procedure can be achieved by implementing a pressure correction based on the average velocity defect w at each cross section of the W , where W is the average velocity at channel such that w = W i a given cross section and Wi is the average inlet velocity. Assuming that the local velocity defect is associated with a pressure defect, the following equation can be formed: Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

Fig. 2 Grid distribution in the xy plane

Fig. 4 Centerline velocity and pressure drop in pipe ow

w p = t z

Thus, the pressure eld is updated with the above correction, and the momentum equations are solved again with the new pressure eld. This procedure dramatically improves the computational efciency 26. Figure 2 shows a typical grid distribution used in the present work. The grid points are clustered near the channel walls and in the entrance region where large gradients exist. The effect of grid resolution on numerical accuracy was studied by performing computations at Re= 100 for a channel with = 45 deg, and the results are given in Table 1 and Fig. 3. Table 1 shows that the variation in the fully developed friction factor is 1% for the last three grids, and Fig. 3 shows that a grid density of 121 provides satisfactory resolution in the axial direction. Therefore, a mesh size of 51 51 121 x , y , z was chosen with an expansion ratio of 1.06 in all directions. However, the grid density in the y direction was

Table 1 Grid size effect on fully developed Pofd = f Refd at different aspect ratios for trapezoidal duct with = 45 deg

increased to 61 with an expansion ratio of 1.07 for = 2 to maintain the accuracy. For a given geometry, the axial extent of the computation domain was set to a minimum of twice the entrance length estimated based on the prevailing Re, which ensures that fully developed conditions are reached well before the channel exit. Extensive code validation studies have been conducted for benchmark ows through circular and rectangular ducts. Because of space limitations, only two examples are presented here. Figure 4 shows a comparison of pressure drop and centerline velocity with data from Shah and London 1 for pipe ow. The results show reasonably good agreement. The small discrepancy between the velocity results close to the inlet can be attributed to the fact that their data are based on some linearization in the numerical scheme, which prevents the solution from predicting the velocity overshoots near the pipe wall. Such velocity overshoots close to the wall lead to lower velocities in the central region due to conservation of mass. Figure 5 compares the calculated pressure drop in the main ow direction of a square duct with the experimental results of Beavers et al. 27. The experimental data are provided for Reynolds numbers in the range 651 Re 1858. Present calculation at Re = 1500 shows a reasonably good agreement with the experimental data. The numerical results of Han 8, and Wiginton and Dalton 28 are also included in the same gure for comparison. Furthermore, in Tables 2 and 3, different ow parameters and heat transfer rates for fully developed ow in trapezoidal ducts with various side angles and aspect ratios are compared to those of Shah and London 1. The overall agreement is found to be very satisfactory, such that the differences are typically 1.5%.

Fig. 3 Axial grid resolution

Fig. 5 Axial pressure distribution in a square duct

Journal of Heat Transfer

JANUARY 2006, Vol. 128 / 65

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

Table 2 Comparison of fully developed ow parameters for trapezoidal channels at different aspect ratios and side angles with Shah and London 1

Results and Discussion

It is benecial to start with the global features of the ow development in the entrance region by considering the velocity elds at different axial locations along the channel. In the fully developed region, nonaxial velocity components are zero and pressure is uniform over a given cross section. However, in the entrance region, the ow is three-dimensional with a nonuniform pressure eld at each cross section. Figures 6a6c show the velocity vectors and pressure contours at three axial planes z+ = 5.21 105 , 6.83 104 , 1.50 103 in a trapezoidal channel with = 30 deg and = 1 at a Reynolds number of 100. The nondimensional axial length is dened as z+ = z / DhRe, and W2 i / 2 is used to normalize the pressure. In Fig. 6a, which is very close to the inlet, a strong secondary ow has formed toward the core of the channel. This ow is the result of large pressure gradients that develop near the walls because of the no-slip boundary condition coupled with the prescribed uniform inlet velocity prole. Note that pressure is maximum locally just inside the acute side angle . Farther downstream Fig. 6b, the ow is more developed in the core of the channel; however, there are still large pressure gradients near the walls. The ow eld shown in Fig. 6c is at a plane closer to the fully developed region as evidenced by the much weaker pressure variation over the cross section: p = pmax pmin = 3 as compared to p = 21 in Fig. 6a. Strong pressure gradients that exist near the walls not only force the ow toward the core of the channel but also push the uid near the walls in the axial direction. As shown in Fig. 7, this process causes the maximum velocities to be found near the walls rather than the channel core as would normally be expected. These unexpected velocity overshoots have been observed by other researchers as well e.g., 1,29. It should be noted that, in order to capture this phenomenon, it is necessary to include both the axial
Table 3 Comparison of the fully developed Nu at different aspect ratios and side angle = 90 deg.

diffusion and transverse pressure gradients in the analysis. The velocity proles in Fig. 7 appear to be symmetric with respect to the centerline of the cross section; however, for ows in channels with higher aspect ratios or larger side angles, asymmetry is clearer. Velocity overshoots are more pronounced near the corners of channels because the pressure gradients are the highest at these locations. For example, in a trapezoidal duct, the region of the acute side angle contains the strongest overshoot, as clearly seen in Fig. 8. In this gure, the 3D axial velocity proles at four axial locations are shown for the same geometry and Re as those in Figs. 5 and 6. The velocity contours in Figs. 8a8c show

Fig. 6 Velocity vectors and pressure contours in the xy plane at three axial locations: a z+ = 5.21 105, b z+ = 6.83 104, and c z+ = 1.50 103

66 / Vol. 128, JANUARY 2006

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

Fig. 7 Axial velocity proles along the y = 0 plane

clearly that the highest axial velocities are located in the sideangle region. The slight reduction in axial velocity in the core of the channel, observed in Fig. 8a very close to the inlet, is simply due to conservation of mass. Another feature of the velocity proles in Figs. 8a and 8b is the presence of very large velocity gradients near the walls, resulting in correspondingly large friction coefcients early in the entrance region. Figure 8d shows the fully developed velocity prole. Very low velocities and gradients in the acute angle region are the notable features in this gure. Channels with other aspect ratios and side angles develop similar ow patterns in the entrance region. Even in the limiting case of rectangular cross sections as the side angle becomes 90 deg, the maximum velocities occur in the corner regions. An example is given in Fig. 9 for a square duct. In the following, the effects of and on the ow patterns and on the key engineering parameters friction factor and heat transfer coefcient will be discussed. Furthermore, axial variations of ow parameters, such as the momentum rate and kinetic energy both closely associated with the ow development in the entrance region, are examined. Four side angles = 30, 45, 60, 90 deg and three aspect ratios = 0.5, 1 , 2 will be considered. The Reynolds number associated with all cases discussed in this section is 100. The effects of Re on the ow features will be discussed later. One of the key parameters of interest in the entrance region is the pressure drop in the principal ow direction. The present analysis indicates pressure variations in both the axial and the transverse directions in the entrance region. The transverse pressure distribution is a function of axial location and Reynolds number. As the ow approaches the fully developed region, the pressure variations in the transverse direction diminish and a uniform pressure over a given cross section is attained for all Reynolds numbers. Because of the presence of transverse pressure variations in the entrance region, a local mean pressure p is dened as p=

p dA /A
i i

Fig. 8 Three-dimensional velocity proles at four axial locations: a z+ = 4.39 104, b z+ = 2.18 103, c z+ = 8.91 103, and d z+ = 5.26 101

where pi is the local pressure at each point on the cross-sectional plane, dAi is the area of the associated control volume face, and A is the cross-sectional area of the channel. The nondimensional pressure drop along the channel is traditionally expressed in the form of an apparent friction factor dened as 4z p = f app = 4z+ f appRe = 4z+Poapp Wi2/2 Dh 6

where p indicates the pressure drop from the entrance, that is from z = 0, and Po is the Poiseuille number. In Fig. 10, the variation of f appRe is shown at Re= 100 for a trapezoidal channel with = 45 deg and three different aspect raJournal of Heat Transfer

tios. For the case with = 1, the circumferentially averaged local skin friction factor f Re= Rew / W2 i / 2 is also included for comparison. Unlike the fully developed region where the pressure drop is only due to the friction at the wall, in the developing region the change in momentum rate accounts for a major part of the pressure drop, as clearly seen in Fig. 10. Obviously, the apparent friction factor approaches the skin friction factor as the ow becomes fully developed. Figure 10 also shows that decreasing the aspect ratio from 2 to 1 has a minor effect on the f Re fd, but decreasing further to 0.5 results in an increase of about 12% in f Re fd. This can be explained by considering the fact that, for JANUARY 2006, Vol. 128 / 67

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

Fig. 10 Variation of apparent friction factor along the axial direction for different aspect ratios

Fig. 11 Variation of hydraulic diameter arbitrary units with channel aspect ratio and side angle

velopment of the ow eld. The momentum rate in the main ow direction is given by KdW2 i A, where Kdz is a momentum ux correction factor dened as 1
Fig. 9 Velocity contours for a square channel at two axial locations: a entrance region z+ = 2.52 103 and b fully developed region z+ = 6.07 101

Kd =

1 A

w/Wi2dA

a given , the hydraulic diameter increases signicantly with increasing aspect ratio, as shown in Fig. 11. At a given Re, this leads to a higher inlet velocity in a channel with lower aspect ratio and, therefore, higher friction factor. In Fig. 12, the variation of the apparent friction factor along channels having the same aspect ratio = 1 but different side angles is shown. It appears that the apparent friction factor is a weak function of for side angles 45 deg. However, at lower side angles, friction is higher in the region closer to the channel inlet but lower in the fully developed region. For clarity, consider a comparison of the = 30 deg and = 90 deg cases. As discussed earlier in relation to Fig. 8, near the channel inlet, the maximum axial velocity is located within the side-angle region Figs. 8a and 8b. The pressure drop required to transform this velocity prole to a fully developed one Fig. 8d is higher than that for a channel with = 90 deg, where the velocity prole undergoes a smoother evolution. However, the fully developed value of the friction factor for = 30 deg is lower due to a higher Dh as explained before Fig. 11. Momentum rate and kinetic energy are also of importance in the entrance region since they are both closely related to the de68 / Vol. 128, JANUARY 2006

Similarly, the kinetic energy of the uid in the main ow direction is given by KeW3 i A / 2, where Kez is a kinetic energy correction factor dened as

Fig. 12 Variation of apparent friction factor in the ow direction for different side angles

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

Fig. 13 Variations of the momentum ux and kinetic energy correction factors along the axial ow direction for different side angles

Fig. 14 Variations of the momentum ux and kinetic energy correction factors along the axial ow direction for different aspect ratios

Ke =

1 A

w/Wi3dA

Figure 13 shows the variations of Kdz and Kez at Re= 100 for a channel with = 1 and different side angles . For a uniform inlet velocity prole, Kd and Ke are expected to increase monotonically from unity at z = 0 to their asymptotic values, Kd and Ke, in the fully developed region. The asymptotic trend observed in this gure is consistent with the fact that maximum velocities increase with decreasing at a xed channel aspect ratio see Table 2. However, near the inlet, this trend is reversed. As it is apparent from Eqs. 7 and 8, the deviations in Kd and Ke from unity are dependent on the amount of ow area over which w / Wi 1, and near the inlet, higher velocities are found close to the walls and particularly in the side angle region. Figure 14 shows the variations of the Kdz and Kez in the ow direction for a given side angle, = 45 deg, while the aspect ratio is varied. In the fully developed region, Table 2 shows that the maximum velocity increases as the aspect ratio increases, and therefore, it is expected that the asymptotic values of Kd and Ke follow the same trend. However, Kd and Ke in the inlet region are not noticeably affected by aspect ratio changes because, near the inlet, the velocity prole development is only weakly dependent on . An approximate analytical method has been developed by Shah and London 1 to determine the incremental pressure drop in the fully developed region, K, for ducts of arbitrary cross section using Kd and Ke such that K = 2Ke Kd 9 For long ducts, total pressure drop can then be estimated based on the knowledge of fully developed friction factor and K using Po PL

Wi2/2

= f Re fd

4L + K DhRe

10

where Po is the inlet pressure and PL is the pressure at a distance L from the inlet. In Table 2, the calculated K are listed for different aspect ratios and side angles, which show good agreement with the results of Shah and London 1. Journal of Heat Transfer

4.1 Thermal Entrance Region. Simlar to velocity, nondimensional uid temperature, = Tw T / Tw Tm, undergoes a continuous variation from the prescribed uniform inlet value = 1 to a fully developed prole such that / z = 0. Here, Tw is the specied constant wall temperature, and Tmz is the mean uid temperature. Figures 15a15d show the 3D temprature contours at four different axial locations corresponding to the velocity proles in Figs. 8a8d. Each gure identies ten isotherms uniformly distributed between the minimum temperature at the wall and the maximum temperature at that location. Similar to the velocity prole, close to the inlet at z+ = 4.39 104 in Fig. 15a, the temperature prole also shows strong gradients near the walls, whereas the central region of the cross section is almost unaffected by the wall temperature. Obviously, unlike the velocity prole, which develops overshoots near the walls and especially within the side angle area see Fig. 8a, overshoots cannot appear in the temperature prole. Under the assumption of constant wall temperature, Tw Ti, and negligible viscous dissipation, heat can only diffuse from the wall toward the center of each cross section with the maximum temperature at the wall. Clearly, velocity overshoots near the walls close to the inlet lead to the higher temperature gradients near the walls. In Fig. 15b at the axial location z+ = 2.18 103, the temperature around the central region is nearly uniform, yet, higher than that in Fig. 15a, with still high temperature gradients close to the wall regions all around the cross section. Closer to the fully developed region at the axial location of z+ = 8.91 103 in Fig. 14c, the temperature prole varies smoothly from = 0 at the wall to its maximum of about = 1.42 in the core. In the fully developed region in Fig. 15d, the uniform wall temperature extends around the side angle areas and the temperature gradients have moved to the central part of the cross section as compared to the temperature proles closer to the inlet Figs. 15a and 15b. This rather large area of almost zero temperature gradient around each side angle accounts for the major reduction in the heat transfer rate for this geometry as compared to the heat transfer rates for larger side angles, where the temperature gradients extend more into the side angle regions. Despite the constant wall temperature, the heat transfer rate is JANUARY 2006, Vol. 128 / 69

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

Fig. 16 Variation of Nusselt number in the main ow direction for different side angles and three aspect ratios: a = 0.5, b = 1, and c = 2

Tm =

1 AWi

n dA TV

12

Fig. 15 Three-dimensional temperature proles at four axial locations: a z+ = 4.39 104, b z+ = 2.18 103, c z+ = 8.91 103, and d z+ = 5.26 101

neither peripherally nor axially constant in the entrance region. The circumferentially averaged local heat transfer coefcient or the local Nusselt number is dened as Nuz = hzDh DhT/nwz = k Tw Tm 11

where T / nw is the peripherally averaged temperature gradient at the wall at any axial location, and Tmz is the mean uid temperature dened as 70 / Vol. 128, JANUARY 2006

In the present work, Pr= 1 in all cases considered, and there+ = z / DhRePr or the recipfore, the nondimensional axial length zT rocal Graetz number commonly encountered in problems of this type is the same as z+ dened earlier. The variations in the heat transfer rate along the main ow direction for channels with three different aspect ratios are shown in Figs. 16a16c for various side angles. It is interesting to note that close to the inlet where strong temperature gradients develop very close to the wall see Figs. 15a and 15b, the Nusselt number is almost independent of . However, as the fully developed region is approached and the temperature prole develops throughout the cross section, the geometry plays an important role on the heat transfer rates. In this region, the effect of the side angle on Nu is particularly strong, such that increasing for a given aspect ratio leads to substantial increases in Nu. On the other hand, heat transfer rates decrease with increasing aspect ratio at xed . Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

Fig. 18 Axial variation of the apparent friction factor with Reynolds number

appropriate for gases. It should be noted that the ow development becomes independent of the Reynolds number if boundarylayer type of assumptions axial diffusion and cross ow pressure gradients ignored are adopted, which was done in most of the prior studies of the current problem. Figures 18 and 19 show the axial variations of Po= f appRe and Nu as a function of Re for a trapezoidal channel with = 1 and = 45 deg. Clearly, the values of Po and Nu in the entrance region are much larger than those in the fully developed region, which highlights the critical importance of the entrance region in determining the uid ow and heat transfer characteristics of short channels found in many modern miniaturized devices. Figures 18 and 19 indicate that close to the inlet where the axial gradients are considerable, the effects of Re cannot be ignored. As expected, the effects of Re extend to higher axial positions at lower Re as axial diffusion gains more weight. These gures also show that the effects of Re become small only for z+ over 0.05 and Re over 50, approximately. Although not shown here, similar trends are observed in other channel geometries. 4.3 The Entrance Length. Starting with a uniform velocity prole at the channel inlet, the hydrodynamic entrance length L is dened as the axial location where maximum velocity attains 99% of its fully developed value. In one of the earlier studies, McComas 30 developed the following entrance length correlation, which employs data associated with the fully developed region such that L+ = Wmax/Wi2 1 K L = DhRe 4 f Re fd 13

Fig. 17 Isotherms for various aspect ratios at axial location z+ = 1 102 for ow at Re= 100 and = 45 deg: a = 0.5, b = 1, and c = 2

Since the geometry of the channel is the same for = 0.5 and = 2 when = 90 deg, the corresponding axial variations in Nu are the same in Figs. 16a and 16c. Therefore, their comparison to those of other side angles presents a measure of the reduction in the heat transfer rates due to an increase in aspect ratio. In addition, comparison of isotherms at different aspect ratios can provide some insight about its effects on the heat transfer rates. Figures 17a17c show a set of identical isotherms in a cross section located at z+ = 0.01 for = 45 deg with three different aspect ratios. Clearly, the temperature gradients next to to the channel walls are stronger for cross sections with smaller aspect ratios. Also, note that the region in the vicinity of the side angle is very poor in heat transfer due to the presence of weak temperature gradients. For a given aspect ratio, this region extends in size as the side angle is reduced and causes substantial reductions in heat transfer as indicated earlier in Fig. 15d. In Table 3, the fully developed values Nu for rectangular ducts are compared to published results 1,33 for different , and good agreement is observed. 4.2 Effects of Reynolds Number. In developing channel ows, the apparent friction factor is a function of duct geometry, boundary conditions, Reynolds number, and axial position. For heat transfer, the Prandtl number of the uid also plays an important role. However, the present work is limited to Pr= 1, which is Journal of Heat Transfer

In Table 4, L+ values for trapezoidal ducts with different aspect ratios and side angles are given for Re= 100. Entrance lengths

Fig. 19 Axial variation of Nusselt number with Reynolds number

JANUARY 2006, Vol. 128 / 71

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

Table 4 Variation of the entrance length L+, for trapezoidal ducts with different aspect ratios and side angles at Re= 100

calculated using Eq. 13 are also tabulated, which are almost identical with the results provided by Shah 14 using the same correlation. It is seen that in all cases, L+ predictions based on the McComas correlation are about three times smaller than the values obtained in the present work. It should also be noted that this correlation does not yield accurate results for rectangular ducts either. For example, L+ = 0.0255 calculated by Shah 14 using Eq. 13 for a rectangular duct with = 0.5 is much smaller than L+ = 0.085 calculated by Wiginton and Dalton 31 and also in the present work see Table 4. Clearly, the McComas correlation is highly inaccurate and should not be used for entrance-length determination. Present data in Table 4 indicates that the entrance length is a relatively weak function of the aspect ratio for a specied side angle . However, for a given aspect ratio, the entrance length increases substantially with decreasing because vorticity generated at the walls must then spread into a larger ow area before fully developed conditions can be attained. In Table 5, the effects of Re on the entrance length are shown for a duct with = 1 and = 45 deg. As expected, the entrance length increases substantially with increasing Reynolds number from L 1.7Dh at Re= 10 to 11Dh at Re= 100, and 55Dh at Re = 500. However, consistent with the results of the previous section, the L+ approaches essentially a constant value as Re increases beyond 100.

Fig. 20 Comparison of available data to the proposed friction coefcient correlation

f Re fd = 13.9


90

0.07

+ 10.4 exp 3.25


90
0.23

14 Concerning the aspect ratio , it should be noted that ows in rectangular ducts = 90 deg are invariant to a 90 deg rotation of the cross section; that is, a channel with the reciprocal aspect ratio has exactly the same ow features. This ambiguity is overcome simply by dening as the ratio of the short side to the long side of the duct, and thus, the aspect ratio for rectangular channels only always remains in the range 0 1. In heat transfer, the Reynolds number dependence is more pronounced at low Re, where axial conduction in the uid cannot be readily ignored. The relative importance of axial conduction was examined by Pahor and Strnad 32 for ow between parallel plates with constant wall temperatures. They showed that axial conduction increases the heat transfer coefcient such that Nu 8.117 for Pe1, which represents a signicant increase as compared to the other asymptotic limit of Nu= 7.541 for high Pe ows. This effect has been incorporated into the proposed heat transfer correlation given below for the range 0.1 Re 1000. Available heat transfers data for the same side angles and aspect ratios as those of Eq. 14 are correlated with Eq. 15 within 8% as shown in Fig. 21. Nu fd = 2.87

Correlations

For fully developed ows, the friction and heat transfer coefcients are functions of the duct geometry and Reynolds number. The present correlation for the friction factor, Eq. 14, is developed based on available data for sides angles = 30, 45, 60, 90 deg and covers the full range of aspect ratios ranging from = 0 for parallel plates to for ducts with triangular cross sections. Data for geometries not considered in the present study are taken from Shah and London 1. Equation 14 predicts available data within 3% as shown in Fig. 20.


90

0.26

+ 4.8 exp 3.9


90
0.21

G 15

G = 1 + 0.0751 + exp 0.45Re

16

Finally, the entrance lengths calculated in the present study are correlated with Eq. 17 as shown in Fig. 22. With this correlation, data for the same side angles as above but for a limited range of aspect ratios 0.5 2 are predicted within 15% for 10 Re 1000 0.8 L = 0.085Re + 0.3 Re Dh


90
0.6

1 + 0.24

17

Table 5 Variation of the entrance length L+, for a trapezoidal duct with = 1 and = 45 deg as a function of Re

Conclusions

Simultaneously developing 3D laminar ow and heat transfer in the entrance region of trapezoidal channels have been studied for 10 Re 1000 and Pr= 1. Because of the existence of high shear 72 / Vol. 128, JANUARY 2006 Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

are such that larger aspect ratios or smaller side angles both result in longer distances from the walls to the core of the channel. Therefore, the vorticity diffusion time is increased, which leads to longer entrance lengths. Furthermore, approximate methods based on data associated with the fully developed region of the channel fail to predict the entrance length properly for trapezoidal ducts. Based on the present solutions of the full 3D governing equations, new engineering correlations are proposed for the entrance length, as well as the friction and heat transfer coefcients for trapezoidal channels.

Acknowledgment
This work was made possible through the nancial support of the Natural Sciences and Engineering Research Council of Canada.

References
1 Shah, R. K., and London, A. L., 1978, Laminar Flow Forced Convection in Ducts, Academic Press, New York. 2 Shah, R. K., and Bhatti, M. S., 1987, Laminar Convective Heat Transfer in Ducts, Handbook of Single-Phase Convective Heat Transfer, S. Kakac, R. K. Shah, and W. Aung, eds., Wiley, New York, Chap. 3. 3 Schiller, L., 1922, Die Entwicklung der Laminaren Geschwindigkeitsverteilung und Ihre Bedeutung fur Zahigkeitsmessungen, Z. Angew. Math. Mech., 2, pp. 96106. 4 Schlichting, H., 1934, Laminare Kanaleinlaufstromung, Z. Angew. Math. Mech., 14, pp. 368373. 5 Schlichting, H., 1968, Boundary-Layer Theory, 6th English Edition, McGrawHill, New York, pp. 176177. 6 Schmidt, F. W., and Zeldin, B., 1969, Laminar Flows in Inlet Section of Tubs and Ducts, AIChE J., 154, pp. 612614. 7 Langhaar, H. L., 1942, Steady Flow in the Transition Length of a Straight Tube, ASME J. Appl. Mech., 9, pp. A55A58. 8 Han, L. S., 1960, Hydrodynamic Entrance Lengths for Incompressible Laminar Flow in Rectangular Ducts, ASME J. Appl. Mech., 27, pp. 403409. 9 Lundgren, T. S., Sparrow, E. M., and Starr, J. B., 1964, Pressure Drop due to the Entrance Region in Ducts of Arbitrary Cross-Section, ASME J. Basic Eng., 86, pp. 620626. 10 Prakash, C., and Liu, Ye-Di., 1983, Analysis of Laminar Flow and Heat Transfer in the Entrance Region of an Internally Finned Circular Duct, ASME J. Heat Transfer, 107, pp. 8491. 11 Asako, Y., and Faghri, M., 1988, Developing Laminar Flow and Heat Transfer in the Entrance Region of Regular Polygonal Ducts, Int. J. Heat Mass Transfer, 31, pp. 25902593. 12 Asako, Y., Nakamura, M., and Faghri, M., 1988, Three-Dimensional Laminar Heat Transfer and Fluid Flow Characteristics in the Entrance Region of a Rhombic Duct, ASME J. Heat Transfer, 110, pp. 855861. 13 Hwang, G. J., Cheng, Y. C., and Ng, M. L., 1993, Developing Laminar Flow and Heat Transfer in a Square Duct With One-Walled Injection and Suction, Int. J. Heat Mass Transfer, 369, pp. 24292440. 14 Shah, R. K., 1975, Laminar Flow Friction and Forced Convection Heat Transfer in Ducts of Arbitrary Geometry, Int. J. Heat Mass Transfer, 18, pp. 849 862. 15 Flockhart, S. M., and Dhariwal, R. S., 1998, Experimental and Numerical Investigation Into the Flow Characteristics of Channels Etched in 100 Silicon, ASME J. Fluids Eng., 120, pp. 291295. 16 Sadasivam, R., Manglik, R. M., and Jog, M. A., 1999, Fully Developed Force Convection Through Trapezoidal and Hexagonal Ducts, Int. J. Heat Mass Transfer, 42, pp. 43214331. 17 Morini, G. L., 2004, Laminar to Turbulent Flow Transition in Microchannels, Microscale Thermophys. Eng., 8, pp. 1530. 18 Lawal, A., and Mujumdar, A. S., 1984, Forced Convection Heat Transfer to a Power-Law Fluid in Arbitrary Cross-Section Ducts, Can. J. Chem. Eng., 62, pp. 326333. 19 Lawal, A., and Mujumdar, A. S., 1985, Developing Flow and Heat Transfer to Power-Law Fluids in Square, Trapezoidal and Pentagonal Ducts, Int. Commun. Heat Mass Transfer, 12, pp. 2331. 20 Patankar, S. V., and Spalding, D. B., 1971, A Calculation Procedure for Heat, Mass and Momentum Transfer in Three-Dimensional Parabolic Flows, Int. J. Heat Mass Transfer, 15, pp. 17871806. 21 Briley, W. R., 1974, Numerical Methods for Predicting Three-Dimensional Steady Viscous Flow in Ducts, J. Comput. Phys., 14, pp. 828. 22 Farhanieh, B., and Sunden, B., 1991, Three-Dimensional Laminar Flow and Heat Transfer in the Entrance Region of Trapezoidal Ducts, Int. J. Numer. Methods Fluids, 13, pp. 537556. 23 Chorin, H. A., 1968, Numerical Solution of the Navier-Stokes Equations, Math. Comput., 22, pp. 745762. 24 Dwyer, H. A., 1989, Calculation of Droplet Dynamics in High Temperature Environments, Prog. Energy Combust. Sci., 15, pp. 131158. 25 Wigton, L. B., Yu, N. J., and Young, D. P., 1985, GMRES Acceleration of Computational Fluid Dynamics Codes, AIAA Computational Fluid Dynamics

Fig. 21 Comparison of available data to the proposed heat transfer coefcient correlation

and heat transfer rates, the entrance region is particularly important for short channels. One of the interesting features of the ow development from a uniform prole at the inlet is the emergence of velocity overshoots, such that maximum axial velocities appear near the walls and especially in the corners as opposed to the channel core, which become more pronounced with decreasing side angle . Viscous dissipation has been ignored in the present work, and the walls are maintained at constant temperature; therefore, similar overshoots cannot develop in the temperature proles. Boundary-layer-type approximations are not valid close to the inlet. However, for Reynolds numbers over 50, and several hydraulic diameters away from the inlet, the behavior of the key ow parameters, such as the Poiseuille and Nusselt numbers, become practically independent of Re, which indicates that diffusion in the axial direction can be ignored. Hydrodynamic entrance length is a function of both Re and geometry. For a trapezoidal cross section, the geometric effects

Fig. 22 Comparison of the numerical data to the proposed entrance length correlation

Journal of Heat Transfer

JANUARY 2006, Vol. 128 / 73

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

Meeting, Cincinnati, pp. 6774. 26 Dwyer, H. A., Cheer, A. Y., Rutaganira, T., and Shacheraghi, N., 2001, Calculation of Unsteady Flows in Curved Pipes, ASME J. Fluids Eng., 123, pp. 869877. 27 Beavers, G. S., Sparrow, E. M., and Magnuson, R. A., 1970, Experiments on Hydrodynamically Developing Flow in Rectangular Ducts of Arbitrary Aspect Ratio, Int. J. Heat Mass Transfer, 13, pp. 689720. 28 Wiginton, C. L., and Dalton, C., 1968, Incompressible Laminar Flow in the Entrance Region of a Rectangular Duct, 12th Int. Congr. Appl. Mech., Stanford University, Stanford, CA.

29 Karniadakis, G. E., and Beskok, A., 2002, Micro Flows, Fundamentals and Simulation, Springer-Verlag, Berlin, Chap. 4. 30 McComas, S. T., 1967, Hydrodynamic Entrance Lengths for Ducts of Arbitrary Cross Section, ASME J. Basic Eng., 89, pp. 847850. 31 Wiginton, C. L., and Dalton, C., 1970, Incompressible Laminar Flow in the Entrance Region of a Rectangular Duct, ASME J. Appl. Mech., 43, pp. 854856. 32 Pahor, S., and Strnad, J., 1961, A Note on Heat Transfer in Laminar Flow Through a Gap, Appl. Sci. Res., Sect. A, 10, pp. 8184. 33 Aparecido, J. B., and Cotta, R. M., 1990, Thermally Developing Laminar Flow Inside Rectangular Ducts, Int. J. Heat Mass Transfer, 33, pp. 341347.

74 / Vol. 128, JANUARY 2006

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/04/2013 Terms of Use: http://asme.org/terms

You might also like