1. Introduction
Convective heat transfer is one of the main mechanisms in heat transfer for flow in pipes, such as heat exchangers. It could be improved by changing thermophysical properties of working fluids and/or geometrical change in flow or boundary conditions. One of the methods to accomplish this is adding nanoparticles to the base fluid [
1]. This innovative idea was put forth by Maxwell in 1873 [
2], which has generated a great interest for researchers to develop high-performance heat transfer fluids [
3]. The effects of nanoparticles on heat transfer enhancement, however, have produced challenges, regardless of their positive effects [
4]. These challenges could result from parameters that affect heat transfer enhancement, such as volume fractions, aspect ratio, base fluids, and entrance effects, among others, which are still unclear and require more research.
Advanced technology applications that have adopted nanofluids have attracted a wide spectrum of research. Nanofluids could become very significant when acting as smart fluids and applied to biomedical, heating, and automotive, among other, applications [
5]. The development of artificial organs has highlighted the importance of biofluid mechanics, which is a promising field of research for nanofluids, for example, in the blood vessels/respiratory tract/lymphatic and gastrointestinal systems/urinary tract zones [
6,
7]. The design of electronic thermal management systems critically depends on the nanofluid type used [
8] and is considered very important to the electronic cooling industry [
9]. Also, nanofluids were applied to solar thermal collectors with attention to environmental concerns and production costs [
10,
11] and applied to solar energy desalination to overcome the relatively lower efficiency and lower yield of solar stills [
12]. Nanofluids in automobiles have significant opportunities in areas such as radiator coolants, lubricants, brake fluid, and shock absorbers [
13]. The list of applications is long and forces the current work to focus on nanofluid applications to gaseous laminar flow through pipes.
Vishwanadula and Nsofor [
14] reported, experimentally, that the use of nanoparticles caused an increase in the heat transfer performance in pipe flows, and they developed a correlation in terms of the Reynolds number (Re), Prandtl number (Pr), and Nusselt number (Nu). Likewise, Kaur and Gangacharyulu [
15] demonstrated that including a small-volume fraction of aluminum oxide (Al
2O
3) with fluids has enhanced the particle-fluid mixture’s thermal conductivity when compared with those of base fluids like air or water. Furthermore, a study on the convective heat transfer of γ-Al
2O
3 nanoparticles and deionized water flowing through a tube in the laminar flow regime was published by Wen and Ding [
16]. Their results presented that the use of γ-Al
2O
3 nanoparticles significantly enhanced the entrance region convective heat transfer. The reported enhancement was a function of nanoparticle volume fraction and axial distances. These results were confirmed by the study by Vafaei and Wen [
17], which was supported by a detailed literature review. Trivedi and Johansen [
18] worked on forced convective heat transfer in Al
2O
3–air nano-aerosol. They reported that a small particle mass loading is required to increase the value of Nu. Moghadassi et al. [
19] investigated how a horizontal circular tube’s laminar forced convective heat transfer would be affected by nanoparticles by utilizing the computational fluid dynamics (CFD) approach. They showed that the addition of nanoparticles caused an increase in the pressure drop, the friction factor coefficient, and the coefficient of convective heat transfer. Moreover, Kima et al. [
20] studied, using an experimentally validated CFD model, the convective heat transfer and flow characteristics of nanofluids and indicated that the heat transfer coefficient and Nu increased with an increase in the Re value.
The no-slip wall condition is applicable to Newtonian fluids when the Knudsen number (Kn) is smaller than 0.001. According to Maxwell’s kinetic theory [
2], slip flow is observed when the characteristic flow length scale is within the order of the mean free path of the gas molecules. When Kn is less than 0.1 and larger than 0.01, the temperature jump and velocity slip appear, and continuum models remain valid. Therefore, Navier–Stokes equations could still be used to model the laminar flow inside circular pipes [
21]. Based on this approach, Vocale and Morini [
22] utilized the CFD tool to investigate a rarefied, laminar, fully developed gaseous flow within a rhombic microchannel, taking into account temperature jump conditions. They confirmed the dependence of Nu on the microchannel geometry and Kn. Spiga and Vocale [
23] investigated numerically rarefied fluid flowing in a micro-duct with an elliptical cross-section under slip flow. The steady-state, laminar, and fully developed flow under an axial uniform linear heat flux boundary condition was examined. They validated their model against published results from the literature. Their developed numerical code was reported to be capable of obtaining velocities and temperatures of the flowing fluid inside the elliptical cross-sectional pipes. Akbarinia et al. [
24] conducted a numerical simulation to evaluate volume fractions of Al
2O
3 nanoparticles and inlet velocity effects on enhancing heat transfer in slip and no-slip nanofluid flow regimes. Their findings demonstrated that the major improvement in Nu was caused by an increase in the inlet velocity rather than a rise in the volume fraction of nanoparticles. Recently, Siham et al. [
25] utilized Python software to study the rarefied flow with Cu–H
2O nanofluid particles through a rectangular microchannel. They confirmed the significant effect of Re, Kn, and the nanoparticle volume fraction (ϕ) on the heat transfer within the system.
Rovenskaya and Groce [
26] investigated heat transfer in rough wall surface microchannels with a slip flow regime. They considered incompressible gas flow for a range of Kn [0.01–0.1]. They reported that Nu is inversely proportional to Kn for both smooth and rough channels. This was confirmed by Ameel et al.’s [
27] analytical solution of a fully developed laminar flow of gas in microtubes under constant wall heat flux conditions, in which they reported that Nu was inversely dependent on Kn and this dependence increased as a function of temperature jump. Furthermore, Larrode et al. [
28] stressed that neglecting the temperature jump could lead to a misleading conclusion of the increase in Nu as a function of Kn.
Researchers such as Tunc and Bayazitoglu [
29,
30] analytically investigated the heat transfer within microtubes under laminar and fully developed gaseous flows under uniform heat flux boundary conditions. The effects of axial conduction and viscous dissipation within the fluid on the same problem were solved by Jeong and Jeong [
31]. It was reported that axial conduction within the fluid was constant under constant wall heat flux boundary conditions. Consequently, Nu was not affected, and it became independent of the Péclet number (Pe = Re Pr) [
29]. Ramadan [
32] highlighted the significance of shear work in the convective heat transfer of gas flowing in a microtube. He stated that shear work could be as high as the heat transfer via conduction at the upper limits of the slip flow regime. He added that the effect of shear work consists of viscous dissipation and pressure work combined effects. Knupp et al. [
33] investigated the problem of the conjugate heat transfer of gaseous flow inside a microtube under slip flow regime. They stressed the importance of including the effects of axial diffusion and wall conjugation on the value of Nu at a low Péclet number (Pe < 20).
In an earlier work, Alkouz et al. [
34] conducted a numerical study of Al
2O
3–air nanofluid laminar flow in pipes, taking into consideration the velocity slip and temperature jump. A correlation for the average Nu was developed under a constant surface temperature. The study showed that an increase in temperature jump and slip velocity led to a reduction in average Nu. Sun et al. [
35] conducted a numerical study of a laminar gaseous flow through a pipe under constant wall heat flux and slip flow conditions. They linked the resulting Nusselt number with Knudsen and Brinkman numbers at the developing and fully developed regions. In their comprehensive study, the effects of the addition of nanoparticles to the base fluid was not investigated. Moreover, Su et al. [
36] conducted an analytical study on a fully developed laminar slip flow inside an elliptical microchannel with constant wall heat flux conditions. Their fully developed Nusselt number values were corelated to an air-specific heat ratio, as well as Knudson and Prandtl numbers. Their work failed to report effects of nanofluids at the entrance region of a circular pipe.
Although there are many studies that have dealt with heat transfer enhancement through nanofluids in pipes, the use of slip flow in engineering applications is not widely accepted by the academic community. This is due to the small length scale of the slip flow regime compared with the length scale of the flow/system. Therefore, the rarefied flow of gaseous nanofluids under slip boundary conditions at the entrance region is a fundamental study on a rarely addressed limitation of the fluid flow and requires further attention.
Recently, sustainable development goals (SDGs) have dictated that energy efficiency must be increased from its current level to almost double. Enhancing heat transfer within engineering systems directly supports energy efficiency goals by reducing energy consumption and fostering technological innovation across industrial sectors. In this research, a mixture of air and Al2O3 flow at the circular pipe entrance region under slip boundary conditions was investigated. Developing laminar flows with different values of constant wall heat flux were simulated using the CFD approach based on the Ansys-Fluent 2018 software package. The CFD model investigated the effects of Pr, Re, wall heat flux (q″), Kn, aspect ratio (x/Dh—the distance traveled by the nanofluid along the pipe axis to the hydraulic diameter), and ϕ of Al2O3 on the Nusselt number (Nu), velocity ratio (U/Um), and skin friction coefficient presented by (ReCf). Based on the comprehensive CFD simulations conducted, the correlations of Nu, U/Um, and ReCf as functions of other parameters were presented.
3. Results and Discussion
The effects of the Reynolds number (Re = 250, 500, 1000, 1500, or 1750), Al2O3 nanoparticle volume fraction (ϕ = 0, 0.01, 0.03, 0.05, or 0.1) or Prandtl number (Pr = 0.765, 0.759, 0.750, 0.742, or 0.729), Knudsen number (Kn = 0, 0.01, 0.03, 0.05, 0.07, or 0.1), and aspect ratio (x/Dh > 200) on the entrance length (L/Dh), velocity ratio (U/Um), skin friction coefficient (Cf), and Nusselt number (Nu) are presented in this section at the entrance region. The entrance region was decided based on the value of (x/Dh) that corresponded to the fully developed value (U/Um), as defined in Equation (16). Results from within the entrance region were considered in the current study only.
The slip flow was modeled based on Equations (12) and (13), and it was dependent on the value of the mean free path value (λ) that is defined in Equation (11). The temperature value in Equation (11) was dependent on the constant value of wall heat flux (q″) that was provided at the wall boundary condition. Therefore, three different values of q″ = 5, 10, and 50 W/m
2 were studied, and the results are shown in
Figure 3,
Figure 4,
Figure 5,
Figure 6 and
Figure 7. The value of U/U
m as a function of the inverse of the Graetz number (Gz
−1 = x/(D
h Re Pr)) increased with an increase in the axial distance from a value of 1.0 at Gz
−1 = 0.0 to a value of 1.53 at Gz
−1 = 0.07 for q″ = 5 W/m
2, as shown in
Figure 3a. At q″ = 5 or 10 W/m
2, the value of U/U
m was similar, with minimal differences. This meant that the increase in the fluid temperature in Equation (11) was small enough not to create a major effect on the resulting value of λ, which was reflected in the value of slip velocity defined in Equation (12). However, this was not the case for q″ = 50 w/m
2, where the increase in the nanofluid temperature resulted in a noticeable difference in U/U
m in comparison with the other two values. The highest difference occurred at Kn = 0.1 and was less than 1% when compared with the results at q″ = 5 W/m
2.
Figure 3b shows that the slip wall boundary represented by Kn affected the fully developed value of U/U
m by lowering its value from 2.0 at Kn = 0.0 (no-slip wall) to the value of 1.55 at Kn = 0.1 (slip wall). The other values were found to be located in between these values in a nonlinear pattern due to the conservation of mass. Consequently, a constant average velocity was maintained because of the constant thermophysical property assumption. The combination of the continuity and the momentum equations caused the non-linearity in the U/U
m vs. Kn curve. As the slip velocity value increased due to the increase in the Kn value, the fluid velocity value at the wall increased and became higher than zero. Therefore, the maximum velocity value at the pipe centerline would decrease, resulting in a lowered velocity ratio value, as shown in
Figure 4.
At an entrance length of 10 pipe diameters (10D), the differences in the velocity profile were minimal for all investigated wall heat fluxes due to the relatively low increase in the nanofluid temperature, as shown in
Figure 4a. However, the nanofluid temperature began to increase as the fluid traveled along the pipe axis toward the fully developed point due to the heat transferred from the pipe wall. The increase in the temperature is proportional to the value of the wall heat flux. As the q″ value increased, the λ value increased and resulted in a higher slip velocity. Consequently, the nanofluid axial velocity would decrease to fulfill the constant mass flow rate restrictions in accordance with the conservation of mass law, as shown in
Figure 4b.
At constant properties, the skin friction coefficient (C
f) is defined in Equation (19) and is dependent on the variation in wall shear stress, which is correlated to the viscosity, velocity gradient at the wall, and Kn of the slip wall. As shown in
Figure 5, the highest value of (ReC
f) corresponded to the flow with no-slip conditions (Kn = 0.0), which resulted in the highest velocity gradient at the wall. As the value of Kn increased, the slip velocity value increased, resulting in an increased wall velocity and a reduced velocity gradient at the wall. Consequently, this lowered the shear stress value and the skin friction coefficient (C
f) as shown in
Figure 5a. The highest value was C
f = 16 at Kn = 0.0, which was decreased to C
f = 8.9 at Kn = 0.1, as shown in
Figure 5b. However, the effect of increasing the value of q″ from 5 W/m
2 to 50 W/m
2 was significant and resulted in reducing the value of C
f by 4%. This is due to the increase in the λ value, which resulted in an increased slip effect at the wall and a reduced value of shear stress, which was more than expected. The 4% reduction at Kn = 0.1 remains an acceptable variation, and the results from the current simulations remain valid at the entrance region of the circular pipe.
The estimation of Nu based on the current CFD simulation was performed by collecting the (T
w) value from the CFD simulation and calculated nanofluid bulk temperature (T
m) values based on Equation (15). The value of T
m depends on the wall heat flux and nanofluid properties that could be represented by Re and Pr, at which it changed linearly with the axial distance (x/D
h) traveled along the pipe by the nanofluid, as shown in
Figure 6a. The maximum increases in T
m at the simulated entrance region, in comparison with the inlet temperature, were 2.8, 5.4, and 23.3 K for q″ = 5, 10, and 50 W/m
2 and occurred at x/D
h = 92.6, 89.8, and 77.6, respectively. These increased T
m values resulted in an increased λ value, according to Equation (11), and caused an increase in Kn values based on Equation (10). Therefore, the increase in λ or Kn affected the estimation, by CFD, of the required values of T
w needed to maintain a constant value of q″ = 5, 10, or 50 W/m
2. The maximum nonlinear increases in T
w values, in comparison with the inlet temperature, were 4.4, 8.8, and 41.8 K for q″ = 5, 10, and 50 W/m
2 and occurred at x/D
h = 92.6, 89.8, and 77.6, respectively, as shown by
Figure 6b. Consequently, the value of Nu defined in Equation (14) became dependable on the wall heat flux value.
The maximum value of Nu
x at the entrance of the pipe was 4.8 for all simulated q″ values, as shown in
Figure 7a. The value decreased at different rates based on the applied q″ until it reached the fully developed values of 2.64, 2.64, and 2.53 for q″ = 5, 10, and 50 W/m
2, respectively, at Kn = 0.1. This variation in Nu
x values was due to the variation in temperature jump values at the wall, as defined in Equation (13), as a result of the variation in the λ, Kn, and T
w values. However, the variation in the Nu
x of Nu was less than 3%, and it was considered valid to adopt the current approach of defining Kn to simulate the slip flow or temperature jump at the wall boundary condition.
The slip flow at the wall was dominated by Kn, which changed the value of slip velocity, as defined in Equation (12) and shown in
Figure 8. At the no-slip condition (Kn = 0), the U/U
m value reached the well-known maximum value of two for a laminar flow inside a circular pipe at the fully developed flow point. Similar trends were observed for the slip flow at which the U/U
m ratio increased from a value of one to a maximum value at the fully developed flow point. As the Kn value increased, the maximum value of U/U
m was found to decrease, as shown in
Figure 8a. The drop in the U/U
m ratio reached a minimum value of 1.55 at Kn = 0.1 for all values of Re and Pr. Therefore, the effects of Re or Pr on the U/U
m ratio at a constant Kn value were negligible, as shown in
Figure 8b,c.
Figure 9 shows that the slip velocity value increased due to the increase in the Kn value, which caused the nanofluid velocity value at the wall to increase and become higher than 0 m/s. Therefore, the maximum velocity value at the pipe centerline of two would decrease to maintain constant mass flow rates in accordance with the conservation of mass law which resulted in lowering the U/U
m ratio. At Kn = 0.01 and 10 diameters from the inlet, for instance, the value of the velocity at the wall was equal to 0.01 m/s and increased to be equal to 0.39 m/s at Kn = 0.1, as shown in
Figure 9a. The opposite behavior was observed at the centerline velocity, at which the velocity reduced from 0.92 m/s at Kn = 0.01 to 0.77 m/s at Kn = 0.1. The general form of the velocity curves was not parabolic and confirmed that the fully developed flow point is way beyond 10 diameters from the pipe inlet.
Figure 9b shows parabolic nanofluid velocity curves resulting from the state of fully developed flow. When compared with the 10 diameter results, the values of the velocity at the wall were reduced to 0.075 m/s and 0.315 m/s at Kn = 0.01 and Kn = 0.1, respectively. However, the centerline velocity values were increased to 1.263 m/s and 1.019 m/s at Kn = 0.01 and Kn = 0.1, respectively. As a result, the U/U
m ratio increased as the nanofluid traveled downstream from the inlet, and the maximum value that could be reached decreased with the increase in the Kn value.
It is well-known that the hydraulic entrance length (L/D
h) depends on Re proportional manners for pipe flow at no-slip conditions (Kn = 0). The value of (L/D
h), based on current CFD results, was compared with the well-known equation of Shah and London [
40] and shown in
Figure 10a. The results showed the compliance of CFD results with the equation with a high degree of accuracy for all investigated volume fractions of Al
2O
3 nanoparticles. With the already validated CFD results under the slip flow, as shown in
Figure 2a, the effects of the slip flow on the value of (L/D
h) were examined and compared with the no-slip condition results for the same Re and Pr values (Δ(L/D
h) = (L/D
h)
no slip − (L/D
h)
slip), as shown in
Figure 10b. It can be seen that as the Kn value increased, the reduction in the (L/D
h) value increased due to the decreased value of U/U
m for all investigated Re values. Therefore, the distance required for the nanofluid to travel toward the fully developed value of U/U
m was decreased. Consequently, the hydrodynamic entrance length became shorter with the increase in Kn values when compared with the no-slip conditions, starting from a reduction of about 0.8 diameters at Kn = 0.1 for Re = 250 to a maximum value of around five diameters at Kn = 0.1 for Re = 1750. Therefore, the maximum reduction due to the increase in the Kn value was 5 diameters at Re = 1750 and due to the increase in Re value was 5 − 0.8 = 4.2 diameters at Kn = 0.1, as shown in
Figure 10b. Therefore, the presence of a slip wall reduced the hydraulic entrance length with the increase in the Kn and Re values.
The values of C
f obtained from CFD simulations were estimated based on Equation (19) and shown in
Figure 11. The highest C
f value was at the inlet of the pipe due to the highest velocity gradient at the wall, which resulted in the highest wall shear stress. These values decreased as the nanofluid traveled downstream the pipe, causing the velocity gradient to decrease until it reached the fully developed value of 16/Re. As the slip flow became active, with Kn values greater than 0.01 and less than 0.1, the C
f value started to drop below the no-slip condition. The lowest ReC
f value obtained at Kn = 0.1 was around 8.79, and the highest value under slip flow conditions was equal to 14.78 at Kn = 0.01, as shown in
Figure 11a. Similar to the U/U
m results, it has been found that ReC
f was insensitive to the value of x/(D
h Re Pr), resulting from the addition of nanoparticles to the base fluid with different percentages, as shown in
Figure 11b,c. As explained earlier, the change in λ value has to be significant to cause a significant change in Kn that would increase the slip velocity value. The increase in slip velocity would decrease the wall shear stress, resulting in a decrease in the value of C
f.
The addition of nanoparticles to air would change the thermophysical properties of the resulting nanofluid, the value of Pr, and the flow inlet velocity. These changes would result in changing the magnitude of nanofluid bulk temperature (T
m) in accordance with Equation (15) and shown in
Figure 12a. Increasing the nanoparticle volume fraction would result in a decrease in Pr and an increase in the specific heat of the nanofluid (C
p), which increases the ability of the nanofluid to absorb more heat. Therefore, T
m decreased with the decrease in Pr and increased in the flow direction due to the increase in the traveled axial distance. Consequently, the wall temperature (T
w) would decrease to maintain the enforced constant q″ value at the wall boundary condition, as shown in
Figure 12b. The nanofluid temperature at the centerline of the pipe (T
min) was extracted from the CFD simulations directly without any further calculations and followed the same trend of T
m and T
w, as shown in
Figure 12c.
The dual variation in T
m and T
w as a function of Pr resulted in making the dependence of Nu
x on Pr at a constant Kn negligible, as shown in
Figure 13a. A similar trend was found for different Re values, as shown in
Figure 13b. At low Re values, the nanofluid velocity was low, which resulted in a higher T
m slope, based on Equation (15), at a relatively short entrance distance. As the Re value increased, the nanofluid velocity increased and resulted in a low T
m slope at a relatively long entrance distance, as shown in
Figure 13c. The fully developed value of Nu
x was reached at different entrance distances with similar trends. This explains why all simulated Re values followed the same curve of Nu
x vs. (x/(D
h Re Pr)) at a constant Kn, as shown in
Figure 13b. Moreover, the dependence of Nu
x on Kn was more significant than the dependence on Pr or Re, as shown in
Figure 13d. It was confirmed that as Kn increased, the value of Nu
x decreased along the axial distance traveled by the nanofluid. The slip flow and temperature jump effects at the wall were caused by different values of Kn. As a result, they changed both the shear stress values and the T
w values, which resulted in a decrease in the value of Nu
x. At Kn = 0, the fully developed value of Nu
x was equal to 4.36, and the minimum Nu
x value obtained was equal to 2.5 at Kn = 0.1 for all investigated Pr and Re values. Therefore, Kn has negatively affected both the value of Nu
x and the value of the entrance length (L/D
h).
The thermal entrance length (L/D
h)
th was defined as the hydrodynamic entrance length multiplied by Pr, and it was related to the development of the thermal boundary layer inside the pipe. For the current study, the flow was considered thermally fully developed when the temperature ratio (θ), defined in Equation (20), was equal to 0.8. The inlet temperature (T
in) was constant at 300 K, while T
w and T
min were extracted from the CFD simulations. The current approach was validated against the published work of Shah and London [
40] for the no-slip flow cases and shown in
Figure 14a. As can be seen, the CFD results fitted well with the published correlation for all the investigated nanofluid volume fractions under the no-slip condition (Kn = 0). Building upon this conclusion, the variation in (L/D
h)
th was examined and presented in
Figure 14b as (Δ(L/D
h)
th = (L/D
h)
th-slip − (L/D
h)
th-no slip). The increase in Δ(L/D
h)
th with Re, when compared with same conditions at no-slip conditions, was clear, with a maximum value of 14.3 diameters. This is translated to an increase in the thermal entrance length by 14.3 diameters for the slip flow, with Kn = 0.1 at Re = 1750. At Re = 250 and Kn = 0.1, however, the maximum increase was around 2.0 diameters. For the remaining Kn values, the maximum increase in the thermal entrance length was between 2 diameters and 14.3 diameters. Hence, the presence of a slip wall resulted in a proportional increase in the thermal entrance length as both the Kn and Re values increased.
The increase in Kn caused the wall-adjacent cell temperature to decrease due to the temperature jump effect, as defined in Equation (13) and shown in
Figure 15. At the inlet of the pipe, θ was equal to one and decreased as the nanofluid traveled downstream along the axial direction. Both effects required the nanofluid to travel longer downstream distances to achieve the T
min value that satisfied the condition of (θ = 0.8) for a thermally fully developed flow. With the increase in Re, the amount of nanofluid flowing inside the pipe increased, which caused a decrease in T
min and resulted in an increase in the value of (L/D
h)
th.
The enhancement in heat transfer due to the inclusion of nanoparticles (Al
2O
3) into air was represented by values of the local Nusselt number, as defined in Equation (21). The equation presents how much more heat was transferred from the pipe walls to the nanofluid due to the addition of nanoparticles under slip flow conditions when compared with no-slip conditions. Based on the current CFD simulations, the maximum enhancement due to the addition of nanoparticles was found to be 2.5% at Kn = 0.1, Re = 250, and ϕ = 10% or Pr = 0.729, as shown in
Figure 16. The variation in the simulated Pr was around −4.7%, between the maximum volume fraction of nanoparticles within the nanofluid and air. Also, increases in the nanofluid conductivity and viscosity of around 33% and 30%, respectively, were reported. Therefore, the increase in thermal conductivity would enhance the heat transfer rate, while the increase in viscosity would reduce the rate of heat transfer. Furthermore, with regard to the reduction in nanofluid-specific heat due to the addition of nanoparticles, the convective heat transfer tends to increase and cause an overall heat transfer enhancement. Therefore, there was a compromise in the resulted heat transfer enhancement for the simulated values in favor of a slight enhancement.
The effects of slip flow conditions on the local Nusselt number were defined in Equation (22). The equation presents how much less heat was transferred from the pipe walls to the nanofluid due to the increase in the slip wall conditions (increase in Kn value) when compared with no-slip conditions. As the Kn value increased, the resulted slip velocity increased, which increased the convective effect of heat transfer. However, higher Kn values caused higher temperature jump values, which resulted in a significant reduction in heat transfer from the pipe walls to the flowing nanofluid. The results showed a relatively linear degradation in the amount of heat transfer as a function of the Kn value in a proportional manner. As a result, the highest reduction due to slip flow conditions was found to be 40% at Kn = 0.1. This conclusion highlights the importance of Kn over the volume fraction of nanoparticles for slip flow conditions inside a pipe for air and Al
2O
3 nanoparticles.
The comprehensive CFD study conducted has resulted in many data that are suitable for developing a correlation between U/U
m, ReC
f, and Nu
x and the key investigated parameters (x/D
h, Re, Pr, Kn). CurveExpert Professional version 2.7.2. is a reliable software that was used to develop the required nonlinear correlations. The general form of the correlation is defined in Equation (23), and the related constants are listed in
Table 3. These correlations were based on the following ranges of the investigated parameters: 0 ≤ Kn ≤ 0.1, 250 ≤ Re ≤ 1750, 0.729 ≤ Pr ≤ 0.765, 0 ≤ ϕ ≤ 0.1, and 0 ≤ x/(D
h Re Pr) ≤ 0.0785. The resulting correlations were limited to the adopted ranges of the steady-state laminar flow inside a circular pipe under constant wall heat flux. Moreover, the developed correlations were generated to fill the gap in this area and produce a reliable correlation to be used in the field of the slip flow of nanofluid at the entrance region of a circular pipe.