Ting-Yu Lin e-mail: [email protected] Satish G. Kandlikar1 Fellow ASME e-mail: [email protected] Thermal Analysis, Microfluidics, and Fuel Cell Laboratory, Department of Mechanical Engineering, Rochester Institute of Technology, 76 Lomb Memorial Drive, Rochester, NY 14623 Heat Transfer Investigation of Air Flow in Microtubes—Part I: Effects of Heat Loss, Viscous Heating, and Axial Conduction Experiments were conducted to investigate local heat transfer coefficients and flow characteristics of air flow in a 962 lm inner diameter stainless steel microtube (minichannel). The effects of heat loss, axial heat conduction and viscous heating were systematically analyzed. Heat losses during the experiments with gas flow in small diameter tubes vary considerably along the flow length, causing the uncertainties to be very large in the downstream region. Axial heat conduction was found to have a significant effect on heat transfer at low Re. Viscous heating was negligible at low Re, but the effect was found to be significant at higher Re. After accounting for varying heat losses, viscous heating and axial conduction, Nu was found to agree very well with the predictions from conventional heat transfer correlations both in laminar and turbulent flow regions. No early transition to turbulent flow was found in the present study. [DOI: 10.1115/1.4007876] Keywords: heat transfer coefficient, microtube, heat loss, axial heat conduction, viscous heating, minichannel 1 Introduction Heat transfer to gas flow in small diameter tubes is important in a number of MEMS devices. To design these microdevices properly, a fundamental understanding of heat transfer and fluid flow characteristics in small diameter channels is essential. Although there is no reason for heat transfer to be affected by diameter at these scales, large deviations from theory are reported in existing literature on experimental data for gas flow in microtubes. The study is aimed at providing quantitative information on the sources for such discrepancies. Heat transfer research in microchannels was first conducted by Tuckerman and Pease [1] in 1981. They demonstrated that the planar integrated circuit chips can be effectively cooled by a laminar flow of water through microchannels with hydraulic diameters of 86–95 lm. Their work was followed by others who explored the heat transfer and fluid flow characteristics in microchannels. For liquid flow in smooth microchannels, the available literature clearly indicates that friction factors are in good agreement with the conventional (macroscale) correlations [2,3]. A number of studies investigated the heat transfer characteristics of liquid flows in microchannels [3–10]. The earlier studies revealed that the heat transfer coefficients can deviate significantly from the predictions of conventional correlations [7–10] due to experimental uncertainty, while some of the recent studies indicated that the heat transfer can be predicted well by conventional correlations for liquid flow [3–6]. A few studies on gas flow characteristics in microchannels found that due to roughness effects, the friction factor was higher than in conventional correlations [11,12]. Peiyi and Little [11] studied trapezoidal channels with a hydraulic diameter in the range of 45–83 lm, while Tang et al. [12] investigated circular tubes with smooth and rough surfaces. The friction factors in both studies were in good agreement with predictions from conventional correlations for smooth tubes, but were higher than the predictions for rough tubes. Microchannels with hydraulic diameters 1 Corresponding author. Contributed by the Heat Transfer Division of ASME for publication in the JOURNAL OF HEAT TRANSFER. Manuscript received September 7, 2011; final manuscript received May 25, 2012; published online February 11, 2013. Assoc. Editor: Jose L. Lage. Journal of Heat Transfer on the order of 1–10 lm were investigated by a few investigators who found that the friction factors were lower than the predictions from conventional theory due to rarefaction effects [13–18]. Kohl et al. [19] reported that their friction factors were in good agreement with the predictions from the conventional theory for gas flow in 25 lm channel in laminar flow and for 100 lm channel in turbulent flow. A delayed transition to turbulent flow was observed with the transition occurring at a Re of 6000 for smooth tubes. Turner et al. [20] indicated that for hydraulic diameters higher than 4.7 lm and Kn < 0.04, the friction factors agreed well with continuum theory. The friction factors for rough channels with a relative roughness of up to 6%, were also found to be the same in laminar flow. There are very few studies reported in literature on turbulent flow, due to difficulties encountered by the higher pressure drops in the experimental setup design. Similarly, in the case of gas flow in microtubes, there are very few experimental studies reported in literature for heat transfer due to complexities in heat transfer experiments at microscale [21]. Acosta et al. [22] applied heat and mass transfer analogy and studied mass transfer coefficients in narrow channels by applying the limiting current method. Choi et al. [14] and Wu and Little [23] indicate that the heat transfer coefficients in microchannels were significantly higher than predictions from the conventional correlations, and that the Colburn analogy was not valid. The actual reasons for this discrepancy will be the focus of this work. From the reviewed literature, it can be summarized that for liquid flow in microchannels there are no differences between experimental data and the theoretical predictions. With channel diameters greater than 15 lm, friction factors agree well with conventional correlations. Similarly, available data for channel diameters greater than 0.1 mm shows that heat transfer coefficients are also in good agreement with predictions from conventional correlations. For gas flow in microchannels with hydraulic diameters smaller than about 7 lm at atmospheric pressure, friction factors are found to be lower than the conventional theory due to rarefaction effects. Yu et al. [24] reported that their friction factor data of water and gas flows in 52–102 lm channels were similar and lower than predictions. However, Judy et al. [2] indicated that in this diameter range, the data for water were in good agreement C 2013 by ASME Copyright V MARCH 2013, Vol. 135 / 031703-1 Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/15/2014 Terms of Use: http://asme.org/terms with conventional correlations. Since the friction factor data of N2 gas by Yu et al. [24] were the same as that of water, microscale effects were certainly not expected for liquid flow in microchannels within that diameter range [3,4,6,25]. Some other effects, such as heat loss, axial conduction and viscous dissipation, may not be important at macroscale and/or in incompressible flow. However, these effects may become important at microscale, Heat loss is another important consideration in heat transfer at microscale, and not accounting for it correctly is believed to be one of the major reasons for the discrepancies found in experimental data reported in literature. Axial conduction was indicated to have an important effect; although it was not quantitatively assessed and corrected in comparisons presented by previous investigators. Regarding viscous dissipation, accurate formulations are available for laminar flow, although there are no correlations available for turbulent flow. With the above shortcomings in the available literature in mind, the present study is undertaken with the following objectives: (i) Experimentally investigate the heat transfer coefficient of compressible air flow in 962 lm diameter smooth tube. (ii) Carefully evaluate the effects of heat loss, axial conduction and viscous dissipation and compare the results with the theoretical predictions in laminar and turbulent regions. 2 Experimental Setup Figure 1(a) shows a schematic of the present experimental setup. The test facility was an open system with the outlet exposed to atmosphere. A high-pressure air bottle and a pressure regulator capable of regulating pressures up to 20 MPa were used to supply air flow in the tube. Mass and volume flow meters with flow rate ranges of 100, 10, and 1 standard liter per minute, and accurate to within 0.8% reading were used to measure the flow rates. Pressure drop of fluid flowing through the tube was measured by two different working range pressure sensors to increase the accuracy. Pressure sensors were calibrated by a pressure calibrator with an Fig. 2 Schematic drawing of test section assembly accuracy of 52 Pa. The inlet air temperature was measured by resistance temperature detectors (RTD) with an accuracy of 0.1 C. Fine wire 36 gage E type thermocouples were attached on the tube to measure the outside wall temperature. Thermocouples were calibrated to be within an uncertainty of 0.1 C. A high current dc power supply with 0–100 A and 0–20 V was used to supply power to the stainless steel tube. The test section was enclosed in a vacuum chamber to reduce the natural convection heat losses. Flow meters, thermocouples, a pressure transducer and a power supply were interfaced with a LabView program for data acquisition. The inlet pressure was measured by a pressure sensor mounted near the inlet as shown in Fig. 1(b). The outlet was at atmospheric pressure as the flow was discharged into the atmosphere after exiting the test section. A schematic drawing of the test tube assembly enclosed in a vacuum chamber is shown in Fig. 2. The two ends of the test tube were bonded to low thermal conductivity Garolite extensions with epoxy. One end of the test section was connected to a highpressure compression tube fitting for fluid inlet while the other side was exhausted to the atmosphere. Two copper connectors were soldered to the two sides of the stainless steel test tube for power supply wires following the same design as reported by Kandlikar et al. [26]. The test section was heated directly by a dc power supply. Three E-type thermocouples were attached on the outer surface of the tube at uniform spacing to measure the surface temperatures. All thermocouple wires and electric wires inside the vacuum chamber were connected to data acquisition system and power supply through vacuum feed-through national pipe thread fittings. The tube cross-section image taken by an optical microscope and the internal tube surface image taken by a laser confocal scanning microscope are shown in Fig. 3. The average tube diameter is 962.0 lm which is calculated from the tube flow area of several individual cross-section images. The surface roughness was determined from the images to be Ra ¼ 0.8 lm. 3 Data Reduction In the present study, the friction factor and Nusselt number were calculated to obtain the fluid flow and heat transfer performance data for air flow in a microtube. The pressure drop and mass flow rate in the microtube were measured to calculate the friction factor. Neglecting the entrance region effect as the tube diameter to length ratio is quite large (343), the friction factor is calculated from the following equation: f ¼ Dpf Dh qavg L 2G2 (1) where DPf is the pressure drop due to friction and qavg is the average of inlet and outlet fluid densities, and Dh, L, and G are the hydraulic diameter, tube length and mass flux, respectively. In compressible flow, the measured total pressure drop DPmea is the sum of friction pressure drop DPf and acceleration pressure drop DPa Fig. 1 (a) Schematic of the experimental setup for heat transfer and pressure drop investigation in a microtube and (b) 3D drawing of measurement equipment mounting DPmea ¼ DPf þ DPa (2) The acceleration pressure drop DPa is given by 031703-2 / Vol. 135, MARCH 2013 Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/15/2014 Terms of Use: http://asme.org/terms Transactions of the ASME while in turbulent flow with Re higher than 2300 the viscous heating is given by 2 qviscous ¼ plu2m L 7 (9) The input heat to the fluid qin is calculated by the following equation: qin ¼ qmea þ qaxial þ qviscous qloss (10) where qmea is the measured total heat input by the DC power supply and qloss is estimated by a separate experiment as described in the next section. 4 Fig. 3 Test tube cross-section viewed from an optical microscope and the internal surface image obtained from a laser confocal scanning microscope, Ra 5 0.8 lm DPa ¼ G2 1 1 2 q2 q1 (3) where q1 and q2 are the fluid densities at the inlet and outlet sections, respectively. The friction factor can also obtained from the equation for 1D isothermal compressible flow given by [27] Dh P21 P22 P1 (4) 2 ln f¼ L RTG2 P2 where P1 and P2 are the inlet and outlet pressures, respectively. The local heat transfer coefficient and Nusselt number are calculated by the following equations: h¼ q00 Tw;x Tf;x and Nu ¼ hDh kf (5) where Tw,x is the internal tube wall temperature that can be derived from the external wall temperature measurement using a one-dimensional, steady-state heat conduction equation with heat generation [26]. A detailed derivation of this equation is given in Appendix A. Tf,x is the local fluid bulk mean temperature calculated from the energy balance for the compressible flow as shown in Appendix B. The local input heat flux q00 is calculated as follows: q00 ¼ qmea qloss Ah Results 4.1 Heat Loss Estimation. Since the heat input needed during an actual test is quite small, on the order of a few Watts, heat losses constitute a large portion of the heat supplied in the present study. Therefore, it is essential to reduce the heat losses from the system. This was accomplished by enclosing the test section in a vacuum chamber. Heat losses are determined by heating the tube without any fluid flowing through the test section. The tube was heated directly by the power supply and the applied power was quantified as the heat loss corresponding to the difference between the local wall surface and the ambient temperature. Figure 4 shows DT (difference between the wall and the ambient temperature) versus qloss for different levels of vacuum in the chamber. It is seen from Fig. 4 that qloss increases with increasing DT. As expected, qloss decreases with a decrease in the vacuum chamber pressure. At DT ¼ 40 C, the heat loss values at pressures of 760 Torr (atmospheric pressure) and 0.008 Torr are 1.7 W and 0.7 W, respectively. The heat loss values at pressures of 0.012 Torr and 0.008 Torr are the same, which indicates that the heat loss at the lower pressures is due to radiation. Further reduction in pressure did not lower the heat losses. Heat loss is a function of wall temperature as shown in Fig. 4, while the wall temperature during a test with fluid flow is a function of the distance from the entrance. It was therefore essential to calculate the local heat loss separately near each wall thermocouple location. To achieve this, heat loss qloss versus the temperature difference DT at three different locations at distance Lh from the entrance was recorded as shown in Fig. 5. The heat loss estimates in the actual heat transfer experiments were then calculated using this plot. As shown in the figure there is no significant difference in the heat loss estimation at these locations. The heat loss is thus a function of the local temperature alone for all three locations. (6) where qmea, qloss, and Ah are the measured power from the power supply, the heat loss, and the heat transfer area. Axial conduction is calculated by the following equation. Details of the axial conduction estimation are given by Lin and Kandlikar [28] qaxial ¼ kw A Tw;x Tw;1 Lh;x (7) Viscous heating is calculated using the detailed equations given in Appendix C. In laminar flow with Re less than 2300 the viscous heating is given by qviscous ¼ 16plu2m L Journal of Heat Transfer (8) Fig. 4 Comparison of heat losses from the test section at different vacuum levels MARCH 2013, Vol. 135 / 031703-3 Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/15/2014 Terms of Use: http://asme.org/terms Fig. 5 Heat loss versus wall to ambient temperature difference for three test sections with different heating lengths In an actual experiment, since the wall temperature increases along the tube length, the heat losses also increase accordingly. To account for the heat losses during a test at a given location, this varying heat loss needs to be integrated along the tube length. To accomplish this, the test section was divided into three zones and the heat losses were calculated for each zone using the following equation: h i L L h;n h;n1 =2 (11) qloss;n ¼ qloss;DTðn1Þ þ qloss;DTðnÞ Lh;total where qloss, DT(n1) is the heat loss calculated from the local temperature difference DTn1 at location n 1. The heat input to the fluid in the zone located between Lh,n1 and Lh,n is given by Lh;n Lh;n1 qin;n ¼ qin;n1 þ qmea qloss;n (12) Lh;total Figure 6 shows qloss/qin as a function of Re at three different locations along the test section length. It is observed that qloss/qin increases with increasing distance Lh. This is because of the higher local wall temperatures that result in higher heat losses at downstream locations. The qloss/qin decreases with an increase in Re, as the heat losses become a smaller fraction of the heat carried away by the fluid. Fig. 7 Comparison of local net heat flux along the test tube with and without considering heat losses for Re 5 4700 and Re 5 2100 Even though the test tube was uniformly heated by the power supply, a constant heat flux assumption needs to be modified due to the uneven heat loss in the different zones of the test tube. Figure 7 shows the calculated local net heat flux along with the heating length with and without considering heat loss at different Re (4700 and 2100). The net heat flux along the tube is a constant value without considering the heat loss, while the local net heat flux value decreases with increasing Lh due to the heat losses. The local heat flux values without considering heat losses at Re ¼ 4700 and 2100 are 4.01 kW/m2 and 2.00 kW/m2, respectively, while the local net heat flux values considering different heat losses for Lh ¼ 25 mm, 125 mm, and 229 mm at Re ¼ 4700 are 3.81 kW/m2, 3.52 kW/m2 and 3.17 kW/m2, respectively. The corresponding values at Re ¼ 2100 are 1.77 kW/m2, 1.48 kW/m2 and 1.10 kW/m2, respectively. The difference in heat flux values with and without considering heat loss varies from about 5% to 21% at Re ¼ 4700, and 12% to 45% at Re ¼ 2100. These differences are large enough to make them necessary to be accounted for in the calculation for local net wall heat flux. 4.2 Uncertainty Analysis. To assess the accuracy of the measurement, an uncertainty analysis was performed. The calculated parameters such as Nu, Re, G and f are generally denoted as dy. They are functions of variables x1, x2, … … xn. The uncertainty in a parameter is determined as described below y ¼ f ðx1 ; x2 ; … … xn Þ; " 2 2 #1=2 @y @y dx1 þ… … … dxn dy ¼ @x1 @xn Fig. 6 Comparison of heat loss, qloss, normalized by the input heat, qin, as a function of Re at three different heating lengths (13) where dx1, dx2, … … dxn are the uncertainties in the independent variables. The uncertainties in the experimental parameters are listed in Table 1. The uncertainty in diameter was calculated using 14 sets of diameter measurements taken at different locations in the tube sample. A laser confocal scanning microscope was used to measure the internal surface roughness. From these measurements, the average tube diameter was 962 lm and the uncertainty was calculated as 1.28 lm. For high Re conditions, flow rates, input power, and the wall to fluid temperature difference (Tw,x – Tf,x) are high, so the uncertainty is low for these parameters. For low Re conditions, however, the uncertainties are high. Figure 8 shows the calculated Nu uncertainty as a function of Re. As expected, the uncertainty in Nu increases with decreasing Re. In this study, the local wall temperatures were maintained at 031703-4 / Vol. 135, MARCH 2013 Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/15/2014 Terms of Use: http://asme.org/terms Transactions of the ASME Table 1 Uncertainties in the experimental parameters Uncertainty Nu (%) Tf,x ( C) Tw,x ( C) Re (%) d (lm) G (%) m_ (%) DP (%) f (%) (Turbulent, laminar) Lh ¼ 25 mm Lh ¼ 125 mm Lh ¼ 229 mm 5–8, 10–23 4–9, 12–103 6–17, 28–516 0.1–0.2, 0.2–0.8 0.1–0.4, 0.6–3.2 0.2 1–7 1.28 1–7 0.9–6 1–9 3–16 0.2–0.9, 1.4–7.8 Fig. 9 Comparison of friction factor data with theoretical predictions Fig. 8 The variation of uncertainty in Nu as a function of Re at three heating lengths similar values for each flow rate by controlling the input power. This is done mainly to eliminate the effect of variations in heat losses, which depend on the local wall temperatures. The heat input and heat flux values are therefore higher at higher Re. In laminar flow, the heat transfer coefficient is constant and independent of Re, which causes (Tw,x – Tf,x) to increase with heat flux and Re. The low heat flux and (Tw,x – Tf,x) values in the low Re case thus lead to higher uncertainty in Nu. The uncertainty in Nu is also a function of heating location distance Lh and increases proportionately. For large Lh, the wall temperature and heat losses are high, which increases the uncertainty in the derived fluid temperature and Nu. At the location Lh ¼ 25 mm, the Nu uncertainty is less than 23%. At Lh ¼ 125 mm and Lh ¼ 229 mm, the Nu uncertainty is quite high and the experimental data are not reliable. However, in the turbulent flow regime, the Nu uncertainty at the three different locations was less than 20%. 4.3 Friction Factor. Figure 9 shows the comparison of friction factor data derived from Eqs. (1) and (4). There is no significant difference observed in the data derived by the two different equations indicating little influence from rarefaction effects. In the laminar flow, the data are in good agreement with Hagen– Poiseuille flow values, while in the turbulent region, the deviation of experimental data from Blasius and Filonenco equations is less than 10% and is considered as a good agreement with the conventional correlations. Comparing the slopes of the f-Re plots, it is concluded that for Re less than 2000 the flow is laminar and that earlier transition from laminar to turbulent flow is not observed. In the laminar flow, the friction factor is slightly higher than predicted at higher values of Re. This is due to higher entrance region effects. Journal of Heat Transfer Fig. 10 Comparison of experimental data for Nu for 962 diameter tube with available correlations after accounting for heat losses, viscous heating and axial conduction; inner diameter of microtube 5 962 lm and Lh 5 25 mm 4.4 Heat Transfer Coefficient. Figure 10 shows Nu as a function of Re for the 962 lm diameter tube at Lh ¼ 25 mm from the entrance. It is seen that in turbulent flow, the data can be predicted very well by the Petukhov–Gnielinski correlation. In laminar flow, Nu agreeed well with the theoretical laminar fully developed flow value of 4.36. In higher Re range of laminar flow, Nu was slightly higher than the laminar fully developed flow value. This was due to the flow being in the developing region. The entrance length for the tube diameter of 962 lm corresponding to Re ¼ 2000 is about 67 mm, which is significantly higher than the heating length Lh ¼ 25 mm. The transition flow was observed at Re in the range 2000–3000. It should be noted that the effects of heat loss, and axial conduction and viscous heating were accounted for in the data plotted in Fig. 10. Without considering these factors, the data were not well correlated with the conventional correlations. 4.5 Effect of Viscous Heating. Viscous heating effects are usually negligible in macroscale systems or with liquid flow in both microscale and macroscale systems due to the relatively low velocities employed. During gas flow is microscale systems, however, the small hydraulic diameter results in a high velocity, and consequently the thermal energy generated due to viscous heating results in additional heat input to the fluid. Viscous heating MARCH 2013, Vol. 135 / 031703-5 Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/15/2014 Terms of Use: http://asme.org/terms Fig. 11 The comparison of viscous heating effect on Nu in turbulent flow region using laminar flow equation, Eq. (8), and turbulent flow equation, Eq. (9) increases the fluid temperature and the calculated heat transfer is altered accordingly. To accurately evaluate the experimental heat transfer coefficient, it is essential to estimate the viscous heating effect. In laminar flow, viscous heating can be derived by considering the laminar fully developed flow velocity profile and is also available in literature [29]. However, there are no equations available in literature to account for viscous heating effects in turbulent flow. In the present study, viscous heating in turbulent flow was estimated from the proposed model using Eq. (9). Figure 11 shows the Nu obtained by including viscous heating from Eqs. (8) and (9) normalized by the theoretical Nusselt number, Nuth, as a function of Re for different heating lengths. Equation (8) represents the equation for viscous heating in laminar flow, while Eq. (9) represents viscous heating in turbulent flow. At the location Lh ¼ 25 mm, there was no significant difference for the derived Nu by the two equations. This was due to the short heating length and the lower magnitude of viscous heating. However, for all other lengths, using the laminar flow equation, Eq. (8), in the turbulent flow region at higher Re results in significantly higher values of Nu due to overestimation of the viscous effects. Using the turbulent flow equation, Eq. (9), in the turbulent region results in Nu/Nuth to be close to 1 for all lengths. The results shown in Fig. 11 indicate that the derived model in the present study is applicable to estimate viscous heating in turbulent flow. Equation (9) was derived based on a 1/7 power law profile as shown in Appendix C. Although the power exponent is not constant but depends on the Reynolds number, the simple power law equation is used in the present work for ease of calculations. Further refinements in this area are recommended. 4.6 Effect of Heat Loss. The heat losses constitute a large fraction of the heat supplied and significantly affect the experimental uncertainty. Figure 12 shows the comparison of derived Nu data with and without considering heat losses. In turbulent flow, the Nu values evaluated with and without considering heat losses are very close, implying that heat loss effects are negligible. However, in laminar flow, the data without considering heat losses are significantly different from the predictions from the conventional theory. The heat loss effect becomes more prominent in the laminar flow region. Without considering heat losses, the heat input to the fluid was overestimated, along with the heat flux, as illustrated in Fig. 7. Similarly, the local fluid temperature was also overestimated, which caused (Tw,xTf,x) to be lower. From Eq. (5) it is seen that heat transfer coefficient is significantly overestimated at higher q00 and Tf,x values. Fig. 12 Comparison of Nu/Nuth with different heat loss correction equations, Eq. (8) based on laminar flow, and Eq. (9) based on turbulent flow Fig. 13 Effect of axial heat conduction in laminar and turbulent regions and comparison of experimental data with Lin and Kandlikar [28] model 4.7 Effect of Axial Conduction. The axial conduction effect is not negligible in gas flow due to the low heat input and low fluid heat capacity. This effect becomes more significant for smaller diameter tubes, higher thermal conductivity tube materials, and lower RePr product according to the model presented by Lin and Kandlikar [28]. Figure 13 compares the data without correcting for the axial heat conduction effects, and by including them as described by model [28] and Eq. (14). Nuk0 ¼ Nuth 1þ4 1 kw Ah;s Nuth (14) kf Af ðRePrÞ2 where Nuk0 is Nusselt number neglecting the effect of axial heat conduction, Nuth is the theoretical Nusselt number and Ah,s and Af are the cross-sectional areas of the channel wall and the crosssectional area of fluid flow in the channel. It is seen from Fig. 13 that the axial conduction effect was negligible for Re higher than 1000 in the present study. For low Re, the Nuk0 is considerably lower than the conventional theory and the value decreases at lower Re. By comparing the data in Fig. 13 to that in Fig. 10, it is seen that the effects of axial heat conduction are small in the turbulent flow region. In laminar flow, however, the effects of axial conduction are quite large, and the Nu decreases at lower Re if these 031703-6 / Vol. 135, MARCH 2013 Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/15/2014 Terms of Use: http://asme.org/terms Transactions of the ASME Table 2 Re 16968 13297 9652 6788 4561 3081 Wall and fluid temperature variation along test section Tw Lh ¼ 25 mm Tw Lh ¼ 125 mm Tw Lh ¼ 229 mm Tf Lh ¼ 25 mm Tf Lh ¼ 125 mm Tf Lh ¼ 229 mm 30.3 32.1 34.8 36.2 36.6 36.8 38.2 42.3 48.7 52.7 54.1 54.4 45.1 51.3 61.1 67.4 69.6 69.8 26.3 26.9 27.8 28.4 28.7 30.1 34.2 37.0 41.5 44.7 46.0 46.2 42.1 47.1 55.0 60.6 62.7 62.6 effects are not accounted properly. For low Re, the Nu is significantly lower than the conventional laminar fully developed flow theory. It is however, in good agreement with the prediction proposed by the Lin and Kandlikar [28] model. The temperature variation of the microtube at different conditions are listed in Table 2. It shows the magnitude of temperatures gradients along the length, as well as the higher temperatures encountered in the downstream region for different flow conditions. The effect of axial heat conduction in the diameter range of the present study is therefore not negligible, especially for low Re. 5 Conclusions Heat transfer and friction factor of air flow in a smooth 304SS 962 lm inner diameter tube was investigated in this study. The effects of heat loss, axial conduction and viscous heating were analyzed in detail. An uncertainty analysis was performed to evaluate the uncertainty in the data as a function measurement location along the flow length and flow rate. Based on the experimental data the following conclusions were drawn: (1) Heat losses are a function of local wall temperature. Since the wall temperature variation along the tube length increases at lower Re for a given power input, the heat losses are higher in the downstream section. (2) For laminar flow at low Re, the wall temperatures in the downstream region are higher resulting in large heat losses and larger uncertainty, making data in this region unreliable. The experimental Nu uncertainty increases with increasing distance (downstream), and with decreasing Re. (3) Viscous heating effects increase with Re and are not negligible for high Re. A new equation is developed for viscous heating in the turbulent region. (4) The effects of axial heat conduction at low Re are not negligible. The model of Lin and Kandlikar [28] is able to predict this effect well. (5) The heat transfer coefficient and friction factor for air flow in a 962 lm tube can be predicted very well by conventional correlations after accounting for the variable heat losses along the length of the test section, axial conduction and viscous heating effects. Acknowledgment This material is based upon work supported by the National Science Foundation under Award No. CBET-0829038. Any opinion, findings, and conclusions or recommendation expressed in this material are those of the authors and do not necessarily reflect the views of the National Science Foundation. Nomenclature A¼ Af ¼ Ah,s ¼ Dh ¼ f¼ G¼ area, m2 cross-sectional area of fluid flow in channel, m2 cross-sectional area of the channel wall, m2 hydraulic diameter, m fanning friction factor, dimensionless mass flux, kg/m2 s Journal of Heat Transfer h¼ kf ¼ kw ¼ Kn ¼ L¼ Lh,x ¼ Nu ¼ Nuth ¼ P1 ¼ P2 ¼ DP ¼ DPa ¼ DPf ¼ DPmea ¼ qaxial ¼ qin ¼ qloss ¼ qmea ¼ qviscous ¼ q00 ¼ Re ¼ Tw,x ¼ Tf,x ¼ Tf,i ¼ um ¼ heat transfer coefficient, W/m2 C thermal conductivity of liquid, W/m C thermal conductivity of tube, W/m C Knudsen number, dimensionless tube length, m local heating length, m Nusselt number, dimensionless theoretical Nusselt number, dimensionless pressure in the inlet region, Pa pressure in the outlet region, Pa pressure drop, Pa acceleration pressure drop, Pa friction pressure drop, Pa measured pressure drop, Pa axial conduction heat, W input heat to fluid, W heat loss, W measured heat from power supply, W viscous heat, W heat flux, W/m2 Reynolds number, dimensionless local wall temperature, C local fluid temperature, C inlet fluid temperature, C mean fluid velocity, m/s Greek Symbols l¼ q1 ¼ q2 ¼ qavg ¼ viscosity, Ns/m2 fluid density in tube inlet region, kg/m3 fluid density in tube outlet region, kg/m3 averaged fluid density, kg/m3 Appendix A: Internal Tube Wall Temperature Calculation From energy equation, for one-dimensional heat conduction with heat generation q 1d dT r þ ¼0 r dr dr kw (A1) where kw is the conductivity of micro tube and q is the internal heat source generated inside the micro tube by the electric power defined as q¼ q Lh p ro2 ri2 (A2) With the adiabatic condition at the outer wall, the inside temperature is given by q 2 q 2 ro 2 ro ri ro Ln Tw ¼ Tw;o þ 4kw 2kw ri (A3) MARCH 2013, Vol. 135 / 031703-7 Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/15/2014 Terms of Use: http://asme.org/terms Appendix B: Calculation of Local Fluid Temperature U¼ The inlet velocity u1 can be determined from the correlation as follows: 4m_ u1 ¼ q1 pd2 (B1) _ q1 are mass flow rate and inlet fluid density, p respecwhere m, ffiffiffiffiffiffiffiffiffiffi tively. The inlet speed of sound, a1 is expressed as a1 ¼ cRT 1 , where c, R, and T1 are ratio of specific heats, gas constant and inlet fluid temperature. Hence, inlet flow Mach number can be determined M1 ¼ ðu1 =a1 Þ and T01 c1 2 M1 ¼1þ 2 T1 (B2) where T01 is the temperature which the fluid element achieves when u ¼ 0, defined as total pressure. Hence, T01 is obtained as c1 2 (B3) M1 T01 ¼ T1 1 þ 2 From the energy balance T02 ¼ q_ þ T01 cp (B4) _ where q_ is heat added per unit mass defined as q_ ¼ ðqin =mÞ. Therefore T01 ðc þ 1ÞM12 ¼ ½2 þ ðc 1ÞM12 T0 ð1 þ cM12 Þ2 (B5) T02 T02 T01 ¼ T0 T01 T0 (B6) From the compressible flow table, local Ma M2 can be determined. The local fluid temperature T2 can be determined from the following equation where T =T1 can be determined from M1: T2 1þc 2 2 ¼ M2 T 1 þ cM22 T2 ¼ T2 T T1 T T1 (B7) (B8) Appendix C: Viscous Dissipation ( 2 2 2 1 @vh @w 1 @vh 1 @w 2 þ þ þ vr þ r @h @z 2 @z r @h 2 ) 2 1 @w @vr 1 1 @vr @ vh þ þ þr (C1) þ 2 @r 2 r @h @r r @z U ¼ 2l @vr @r Assume steady and @w 0;ð@=@zÞ 0 uniform flow, vh 0; ð@=@hÞ 0; 2 @vr U ¼ 2l @r (C2) r 2 uðrÞ ¼ 2u 1 R (C3) @u r ¼ 4u 2 @r R (C4) In laminar flow ðL ðR 0 0 r2 2l 16u2 4 2prdrdx R (C5) U ¼ 16plu2 L (C6) u r 1=7 ¼ um R (C7) In turbulent flow @u 1 1 r 6=7 ¼ um @r 7 R R ðL ðR 1 2 1 r 12=7 2prdrdx um 2 2l U¼ 49 R R 0 0 2 U ¼ plu2m L 7 (C8) (C9) (C10) References [1] Tuckerman, D. B., and Pease, R. F. W., 1981, “High-Performance Heat Sinking for Vlsi,” IEEE Electron Device Lett., 2(5), pp. 126–129. [2] Judy, J., Maynes, D., and Webb, B. W., 2002, “Characterization of Frictional Pressure Drop for Liquid Flows Through Microchannels,” Int. J. Heat Mass Transfer, 45(17), pp. 3477–3489. [3] Yang, C. Y., and Lin, T. Y., 2007, “Heat Transfer Characteristics of Water Flow in Microtubes,” Exp. Therm. Fluid Sci., 32(2), pp. 432–439. [4] Lin, T.-Y., and Yang, C.-Y., 2007, “An Experimental Investigation on Forced Convection Heat Transfer Performance in Micro Tubes by the Method of Liquid Crystal Thermography,” Int. J. Heat Mass Transfer, 50(23–24), pp. 4736–4742. [5] Yen, T.-H., Kasagi, N., and Suzuki, Y., 2003, “Forced Convective Boiling Heat Transfer in Microtubes at Low Mass and Heat Fluxes,” Int. J. Multiphase Flow, 29(12), pp. 1771–1792. [6] Lelea, D., Nishio, S., and Takano, K., 2004, “The Experimental Research on Microtube Heat Transfer and Fluid Flow of Distilled Water,” Int. J. Heat Mass Transfer, 47(12–13), pp. 2817–2830. [7] Peng, X. F., and Wang, B. X., 1993, “Forced Convection and Flow Boiling Heat Transfer for Liquid Flowing Through Microchannels,” Int. J. Heat Mass Transfer, 36(14), pp. 3421–3427. [8] Wang, B. X., and Peng, X. F., 1994, “Experimental Investigation on Liquid Forced-Convection Heat Transfer Through Microchannels,” Int. J. Heat Mass Transfer, 37(SUPPL 1), pp. 73–82. [9] Peng, X. F., and Peterson, G. P., 1996, “Convective Heat Transfer and Flow Friction for Water Flow in Microchannel Structures,” Int. J. Heat Mass Transfer, 39(12), pp. 2599–2608. [10] Adams, T. M., Abdel-Khalik, S. I., Jeter, S. M., and Qureshi, Z. H., 1998, “Experimental Investigation of Single-Phase Forced Convection in Microchannels,” Int. J. Heat Mass Transfer, 41(6–7), pp. 851–857. [11] Peiyi, W., and Little, W. A., 1983, “Measurement of Friction Factors for the Flow of Gases in Very Fine Channels Used for Microminiature Joule-Thomson Refrigerators,” Cryogenics, 23(5), pp. 273–277. [12] Tang, G. H., Li, Z., He, Y. L., and Tao, W. Q., 2007, “Experimental Study of Compressibility, Roughness and Rarefaction Influences on Microchannel Flow,” Int. J. Heat Mass Transfer, 50(11–12), pp. 2282–2295. [13] Pfahler, J., Harley, J., Bau, H. H., and Zemel, J., 1990, “Liquid and Gas Transport in Small Channels,” American Society of Mechanical Engineers, Dynamic Systems and Control Division (Publication) DSC, Vol. 19, pp. 149–157. [14] Choi, S. B., Barron, R. F., and Warrington, R. O., 1991, “Fluid Flow and Heat Transfer in Microtubes,” Proceedings of Winter Annual Meeting of the American Society of Mechanical Engineers, Dec. 1–6, ASME, New York, pp. 123–134. [15] Liu, J., Tai, Y.-C., and Ho, C.-M., 1995, “MEMS for Pressure Distribution Studies of Gaseous Flows in Microchannels,” Proceedings of the IEEE Micro Electro Mechanical Systems, pp. 209–215. [16] Shih, J. C., Ho, C.-M., Liu, J., and Tai, Y.-C., 1996, “Monatomic and Polyatomic Gas Flow Through Uniform Microchannels,” American Society of Mechanical Engineers, Dynamic Systems and Control Division (Publication) DSC, Vol. 59, pp. 197–203. [17] Harley, J. C., Huang, Y. F., Bau, H. H., and Zemel, J. N., 1995, “Gas Flow in Microchannels,” J. Fluid Mech., 284, pp. 257–274. [18] Takuto, A., Kim, M. S., Iwai, H., and Suzuki, K., 2002, “An Experimental Investigation of Gaseous Flow Characteristics in Microchannels,” Microscale Thermophys. Eng., 6, pp. 117–130. [19] Kohl, M. J., Abdel-Khalik, S. I., Jeter, S. M., and Sadowski, D. L., 2005, “An Experimental Investigation of Microchannel Flow With Internal Pressure Measurements,” Int. J. Heat Mass Transfer, 48(8), pp. 1518–1533. [20] Turner, S. E., Sun, H., Faghri, M., and Gregory, O. J., 2001, “Compressible Gas Flow Through Smooth and Rough Microchannels,” Proceedings of 2001 ASME International Mechanical Engineering Congress and Exposition, Nov. 11–16, American Society of Mechanical Engineers, New York, pp. 381–384. 031703-8 / Vol. 135, MARCH 2013 Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/15/2014 Terms of Use: http://asme.org/terms Transactions of the ASME [21] Guo, Z.-Y., and Li, Z.-X., 2003, “Size Effect on Microscale Single-Phase Flow and Heat Transfer,” Int. J. Heat Mass Transfer, 46(1), pp. 149–159. [22] Acosta, R. E., Muller, R. H., and Tobias, C. W., 1985, “Transport Processes in Narrow (Capillary) Channels,” AIChE J., 31(3), pp. 473–482. [23] Wu, P., and Little, W. A., 1984, “Measurement of the Heat Transfer Characteristics of Gas Flow in Fine Channel Heat Exchangers Used for Microminiature Refrigerators,” Cryogenics, 24(8), pp. 415–420. [24] Yu, D., Warrington, R., Barron, R., and Ameel, T., 1995, “Experimental and Theoretical Investigation of Fluid Flow and Heat Transfer in Microtubes,” Proceedings of the 1995 ASME/JSME Thermal Engineering Joint Conference (Part 1 of 4), Mar. 19–24, ASME, New York, pp. 523–530. [25] Steinke, M. E., and Kandlikar, S. G., 2006, “Single-Phase Liquid Friction Factors in Microchannels,” Int. J. Therm. Sci., 45, pp. 1073–1083. Journal of Heat Transfer [26] Kandlikar, S. G., Joshi, S., and Tian, S., 2003, “Effect of Surface Roughness on Heat Transfer and Fluid Flow Characteristics at Low Reynolds Numbers in Small Diameter Tubes,” Heat Transfer Eng., 24(3), pp. 4–16. [27] Choquette, S. F., Faghri, M., Kenyon, E. J., and Sunden, B., 1996, “Compressible Fluid Flow in Micron Sized Channels,” Proceedings of the 1996 31st ASME National Heat Transfer Conference (Part 5 of 8), Aug. 3–6, ASME, New York, pp. 25–32. [28] Lin, T.-Y., and Kandlikar, S. G., 2012, “A Theoretical Model for Axial Heat Conduction Effects During Single-Phase Flow in Microchannels,” ASME J. Heat Transfer, 134(2), p. 020902. [29] Celata, G. P., Cumo, M., McPhail, S., and Zummo, G., 2006, “Characterization of Fluid Dynamic Behaviour and Channel Wall Effects in Microtube,” Int. J. Heat Fluid Flow, 27, pp. 135–143. MARCH 2013, Vol. 135 / 031703-9 Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/15/2014 Terms of Use: http://asme.org/terms
© Copyright 2025 Paperzz