1. Introduction
One of the essential topics in engineering is saving energy. Nowadays, a mechanism to enhance heat removal and energy storage is a critical research topic. The implementation of pin-fins has improved heat removal and reduced the system’s pressure drop. Martin [
1] demonstrated a theoretical approach to predict the performance of chevron-type plate heat exchangers. The furrows of the sinusoidal patterns and the inclination angles of the crests have been proven to be the most vital design parameters concerning fluid friction and heat transfer. Two kinds of flow may exist between the gap between the plates. The first flow involves crossing small sub-streams following the furrows of the first and second plates. This becomes more dominant at lower inclination angles, leading to lower pressure drops. The second flow involves wavy longitudinal flow between two vertical rows of contact points; contrary to the first flow, this prevails at higher inclination angles, resulting in higher pressure drops. The combined effects of the flows are taken to derive an equation for the friction factor as a function of the inclination angle and Reynolds number. This is known as the Lévêque equation, which is utilized for modeling and developing thermal boundary layers in a fully developed laminar or turbulent flow. Applying the Lévêque equation in this study has shown a good agreement in predicting heat transfer coefficients.
Dović et al. [
2] utilized a mathematical model to study the thermal and hydraulic characteristics of a chevron plate heat exchanger with inclination angles of β = 28° and β = 65°. Here, the Reynolds number was in the range of 0.1 < Re < 250. Values of Nu/Pr
1/3 (
μ/
μw)
0.14, which was derived from the Lévêque equation, agree within ±13% of the measured work for channels with inclination angles of β = 28° and β = 65° at 50 < Re < 1400. The results concur with previous literature within discrepancies of ±(15–35)%. The experiment revealed a sudden decrease of the Nu for Re < 30 for a plate with β = 65°. Since the friction factor remains higher than the channel with β = 28°, it favors the use of plates with lower inclination angles for low Re applications.
The thermal performance and pressure drop were investigated experimentally for Al
2O
3-water and MWCNT-water nanofluids in a plate heat exchanger with chevrons at an angle of 60° by Huang et al. [
3]. The results were compared to plain water. The results showed that nanofluids’ heat enhancement was superior to that of water at constant Reynolds numbers, but little change was observed at constant flow velocity. MWCNT-water had more intensive heat transfer deterioration when compared to Al
2O
3-water due to higher viscosity. The pressure drop was found to augment with increasing nanoparticle concentrations; however, at low nanoparticle concentrations, the pressure drop of the nanofluids was similar to plain water. The dimensionless parameter Nu/Pr
0.3 was found to be higher with nanofluids when compared to water. However, as the nanoparticle volume concentrations increased, the Nu/Pr
0.3 decreased. Overall, the heat transfer correlation agreed well with experimental data.
Al-Neama et al. [
4] studied the hydrodynamic and thermal effects of chevron fins in serpentine channels within a heat exchanger. The continuous plate was broken into nine small fins with angles of 30°. The experimental and numerical results demonstrated that the total thermal resistance decreased tediously with the water flow rate. Increasing the Reynolds number increased the pressure drop and the average Nusselt number. This is due to the thermal boundary layer thickness decreasing with higher fluid velocity. The effects of various chevron angles were investigated; as the chevron angle decreased, the total thermal resistance and heat transfer increased.
Contrary to this, as the chevron angle increased, the pressure drop was found to decrease; this is caused by a more significant gap in the secondary microchannel. The friction factor of a plate heat exchanger was computed on CFD simulations using the LES technique by Zhu et al. [
5]. The flow of the working fluid ranged from 10 < Re < 6000, and the inclination angles of the chevrons were 18°, 30°, 38°, 45°, 52°, 60°, and 72°. The objective was to map the relationships between the friction factor, Reynolds numbers, and inclination angles. A friction factor diagram was developed; the friction factor results and the inclination angle over the range of Re resembled the Moody diagram. The inclination angle was analyzed as a roughness element, thus depicting larger inclination angles leading to rougher surfaces. Hence, higher friction factors were observed.
Fernandes et al. [
6] studied a fluid’s fully developed laminar flow in a double-sine chevron plate heat exchanger. Utilizing the POLYFLOW CFD software, the objective was to model the relationships between the tortuosity and Kozeny’s coefficients with the geometric properties of the plate heat exchanger passages. The chevron/corrugation angles of interest were 29.0°, 39.8°, 48.0°, 59.0°, 74.5°, and 84.9°. The plate heat exchanger’s channel aspect ratio and corrugation angles defined the tortuosity coefficient, shape factor, and coefficient pertaining to the
fRe relationships. The tortuosity and the permeability coefficients from the friction factor relationships increased with higher channel aspect ratios and lower chevron angles. The passages of the plate heat exchanger were found to increase with lower chevron angles but experienced little influence from the channel aspect ratio.
The hydrodynamic properties and flow distributions of a plate heat exchanger with two cross-corrugated channels were investigated numerically by Tsai et al. [
7]. The inclination angle was 60°, and the chevron plates were brazed together at an angle of 180°, one on top of the other. The working fluid was water, where the Reynolds numbers ranged from 170 to 1700. The friction factor of the two channels was found to have different trends when comparing numerical to experimental results. There was approximately a 20% deviation when comparing the experimental and numerical pressure drops; this is caused by secondary flows being introduced in the secondary corrugated channels due to complex geometry that was not accounted for by the κ-ε model.
Han et al. [
8] simulated the three-dimensional temperature, velocity, and pressure of a chevron corrugated plate heat exchanger. The κ-ε NRG model was adopted to simulate the hydrodynamic characteristics. Based on the results, the temperature field was found to be very small in the inlet and outlet port positions, while the pressure field was found to decrease along the flow direction. From the velocity field, it was found that no matter whether the fluid inflows or outflows from the ports, there is always a dead-zone corrugated from the side of the port resulting in a low flow rate, causing the inlet temperature to be maintained. Overall, the CFD results were in good agreement with the experimental results.
Kumar et al. [
9] studied the effects of geometrical parameters on the hydraulic and thermal performance of a U-type plate heat exchanger. The plate heat exchanger consisted of two symmetrical plates with chevron angles of 60°/60° and 30°/30°, and one unsymmetrical plate with an angle of 60°/30°. The working fluid is water with Reynolds numbers ranging from 800 to 2300. The objective was to experimentally investigate the relationships between the chevron angle on the mean pressure drop, meaning friction factor, and effectiveness. Higher Reynolds numbers resulted in the momentum overcoming viscous forces. Hence, the viscosity gradient decreased. Additionally, there was less contact of fluid molecules at higher channel velocities, resulting in the friction factor decreasing. Higher chevron angles promoted inner swirling in the inter-plate channels causing the friction factor to rise. The flow maldistribution parameter increased with higher flow rates but decreased with greater chevron angles; this is associated with the variation of overall friction factor resistance, which decreases at higher mass flow rates. Pressure drops were found to increase for higher Reynolds numbers and chevron angles due to higher interferences. Overall, as the chevron angle increased, there was a positive impact on the effectiveness of plate heat exchangers. This is justified by a more significant obstruction in fluid flow, which promotes more turbulence.
The hydraulic performance of a chevron U-type plate heat exchanger was studied analytically for different aspect ratios, Reynolds numbers (200–5800), maldistribution parameters, number of channels, and port sizes by Kumar and Singh [
10]; the working fluid was water. Overall, the analytical results agreed with the experimental results with a ±10% deviation. The friction factor increased with higher Reynolds numbers and with a more significant number of channels. Due to higher maldistribution, the total pressure drop increased with higher mass flow rates, aspect ratios, Reynolds numbers, and more channels. Here, the pressure drop is a function of the number of channels. The flow maldistribution parameter was found to increase when the number of channels increased. However, increasing the aspect ratio decreased the maldistribution parameter since it is a strong function of the hydraulic diameter, the number of channels, and channel velocity. Lastly, the channel velocity increased with larger port diameters due to a significant influence of port diameters on plate heat exchangers.
Naik and Matawala [
11] experimented with the heat transfer of an oil-to-water fluid in a plate heat exchanger. The effects of chevron angles (30°, 45° & 60°) along with other geometric characteristics were investigated to determine their impacts on the heat transfer coefficient. Here, the Reynolds numbers ranged from 50 to 1000, and the Prandtl numbers ranged from 3 to 75. The results depict that as the mass flow rate increases for both oil and water, so does the heat transfer coefficient. Furthermore, increasing the flow rate of water decreased the outlet temperature of the oil; this increase in heat transfer is due to more cooling water being available at higher flow rates. The chevron angle of 60° was found to have the highest heat transfer coefficient. Lastly, increasing the Reynolds number was found to lower the friction factor.
Jain et al. [
12] employed the κ-ε turbulence viscous model for the heat transfer and fluid flow of water in a plate heat exchanger with chevrons at 60° angles. The Reynolds numbers ranged from 400 to 1300, and the Prandtl numbers ranged from 4.4 to 6.3. The results of the numerical study were compared to experimental findings from previous literature. It was found that the experimental friction factors and Nusselt numbers were underpredicted by 2.5–14.5% and 3–18%, respectively. An explanation for this is the exclusion of the port and flow distribution areas in the numerical modeling. Lastly, the velocity vectors were found to inhibit a zig-zag pattern, and the fluid flow is mainly along the main flow direction.
Muley and Mangllik [
13] experimented with a single-phase U-type plate heat exchanger with chevrons to study water’s isothermal pressure drop and heat transfer. The Reynolds and Prandtl numbers ranged from 600 to 10
4 and from 2 to 6, respectively. There are three different types of chevron plate arrangements—30°/30°, 60°/60°, and 30°/60°. The Nusselt number was found to augment with higher chevron angles; this trend is reflected in the intense swirl flow generated by greater chevron angles present in the plates. A higher friction factor was enticed despite the more significant heat transfer coefficients associated with chevrons. The friction factors were 13–44 times greater than flat plate channels due to greater flow friction in the interpolate channel. In addition to the chevron angle, the area enlargement factor influenced the hydrodynamic and thermal characteristics. Higher heat transfer rates and pressure drops were obtained with higher enlargement areas. This is expected as greater swirl mixing is promoted.
The experimental Nu and isothermal data of
f were determined for cooling vegetable oil in a single-plate heat exchanger by Muley et al. [
14]. The Prandtl and Reynolds numbers were in the range of 130 < Pr < 290 and 2 < Re < 400. Three configurations were utilized for the chevron plate arrangements: two symmetrical 30°/30° and 60°/60° plates, and one mixed plate arrangement consisting of 30°/60°. The thermohydraulic performance was heavily influenced by the chevron angle, corrugation aspect ratio, and fluid flow parameters. Under the same flow conditions, the Nusselt number was enhanced up to 3 times in the presence of chevrons compared to the equivalent flat-plate pack, while the friction factor increased 6.6 times. Lastly, depending on the Reynolds number and chevron angle, under fixed surface geometry and pumping power, the heat transfer was enhanced up to 2.9 times more than its flat-plate channels’ counterpart.
Asadi and Khoshkhoo [
15] evaluated the thermohydraulic performances on the water with various chevron angles. The results depicted that the optimal chevron angle was 60°. This was denoted by the highest heat transfer coefficient and the trend for the friction factor. As the chevron angle increases, so does the friction factor; however, at 60°, the results delineate an inverse relationship with the mass flow rate for both laminar and turbulent regimes. Kılıç and İpek [
16] studied the heat transfer rate and overall effectiveness of corrugated plate heat exchangers experimentally. The chevron angles of interest were β = 30° and β = 60°. As depicted by the results, the superior chevron angle was 60°. This is denoted by the higher heat transfer rate and overall effectiveness of the plate heat exchanger. Furthermore, the results also entailed that increasing the Reynolds number resulted in the heat transfer rate and overall effectiveness to augment; this was in agreement with previous studies.
Dolatabadi and Aghdam [
17] presented the experimental and numerical analysis of the heat transfer and fluid flow of a triangular chevron plate heat exchanger. The studies were conducted with air as the working fluid with a uniform heat flux of 1350 W/m
2 applied to the wall. The Reynolds numbers, phase shifts, and channel heights varied in the ranges of 1000 < Re < 10,000, 0° < Ф <180°, 5 mm < D < 35 mm, respectively. Increasing the Reynolds number decreased the fluid temperature significantly, leading to greater Nusselt numbers while simultaneously reducing the friction factor; this subsequently decreased the thermal efficiency factor. The Nusselt number for the chevron channels was higher than the plain channels due to increased turbulence intensities, leading to higher heat transfer rates. The best results obtained when accounting for the Nusselt number and friction factor, i.e., the thermal efficiency index, was at a phase difference of 90°. Lastly, the subsequential increase of distance between the surfaces (D) led to higher temperatures and lower Nusselt numbers/thermal efficiency factors. The volume of vortices causes this in the laminar sublayer, and the fluids’ velocity decreases. As a result, to maximize the efficiency of the heat exchanger, the minimum distance between surfaces should be utilized.
Mohebbi and Veysi [
18] investigated a small, brazed plate heat exchanger’s thermohydraulic characteristics. The chevron angles were 60°, and the modified Wilson plot was incurred to calculate the Nusselt numbers. Chevron corrugations on plates were found to enhance the heat transfer and pressure drop. However, the results depict that the heat transfer and friction coefficients calculated in previous literature disagree up to 34% with the results of this study in some cases. This implies that the correlations presented for large plate heat exchangers cannot be applied for small plate heat exchangers.
A numerical study on the thermohydraulic characteristics of water undergoing turbulent flow between two parallel chevron plates was presented by Jang and Lin [
19]. The flow is incompressible, and the Chen κ-ε model was employed. The effects of three different chevron angles (20°, 40°, and 60°) relative to the direction of the main flow were investigated for inlet velocities ranging from 0.5 m/s–2.0 m/s. Increasing the chevron angle distorts streamlines, leading to stronger vortex zones. Hence, greater Nusselt numbers and pressure drops occur. Nusselt numbers were more significant near the wave bumps due to the wavy configurations interrupting the boundary layer. Lastly, higher inclination angles led to greater Colburn (
j) and friction factors (
f), with the inclination angle of 40° possessing the optimal goodness factor (
j/f).
In the present paper, we investigate the performance of pin-fins in heat removal, forming a chevron shape structure. The novelty of this study is to investigate the presence of Chevron at different heights. Two critical parameters are studied in this analysis. The variable pin-fins height is implemented to investigate whether the higher the pin-fins, the better heat performance. The second is the replacement of the solid pin-fins with porous pin-fins with three different permeabilities of 10 PPI, 20 PPI, and 40 PPI, respectively. The porosity is constant and equal to 0.91 [
20].
Section 2 presents the problem under investigation, followed by the finite element formulation in
Section 3.
Section 4 presents the results and discussion, and finally, the conclusion is in
Section 5.
3. Finite Element Formulation
The full Navier–Stokes equation and the energy equation were solved numerically using the finite element method [
20]. The set of a non-dimensional term used in our analysis are as follows:
Here, De is the hydraulic diameter, kw is the water conductivity, q” is the applied heat flux, Tin is the inlet temperature, and is the inlet velocity. In particular, the momentum equations in the non-dimensional form are as follows in different directions:
Momentum equation along X-direction Here, Re is the Reynolds number set equal to .
Momentum equation along Y-direction Here, V is the velocity component in the Y direction.
Momentum equation along Z-direction Here, W is the velocity component in the Z direction.
Continuity equation
The continuity equation for this simulation can be expressed as
Energy conservation equation
The energy equation is as follows:
Here, is the non-dimensional temperature used in our calculation. The Prandtl number Pr is the ratio of the product of the specific heat of the fluid and the viscosity divided by the fluid conductivity. Thus, Pr is equal to .
3.1. Darcy–Brinkman Model
Since the flow rate within the system is low and the porosity of the metal foam is 0.91, the Darcy–Brinkman equations are used to analyze the fluid behavior. The equations in three dimensions are as follows:
Darcy–Brinkman in X direction Here Da is the Darcy number known to be the ratio of the permeability to the square of the characteristic length.
Darcy–Brinkman in the Y
direction Darcy–Brinkman in the Z direction
The energy equation is written as follows:
Here,
is the porosity set equal to 0.91, and
is the conductivity of the solid porous material. It is important to mention that the conduction equation is used for the solid aluminum. The equation for the conduction heat transfer is shown in Equation (11) as follows:
Here, is the ratio of the conductivity of the aluminum divided by the water conductivity. As the formulation is coupled, the equations were solved simultaneously using the finite element technique.
3.2. Boundary Conditions
At the inlet, the following boundary conditions are applied:
At the outlet, the free boundary condition is applied; thus, the stress is set equal to zero. At the bottom of the plate, as shown in
Figure 1a, the heat flux is set equal to 1. All other surfaces surrounding the model assume zero flux, thus being insulated.
3.3. Mesh Sensitivity Analysis
In order to use the optimum mesh size, mesh sensitivity tests were conducted for the five-chevron configuration.
Table 2 presents the mesh size and the calculated average Nusselt number. Based on the data obtained in
Table 2, the best mesh size is that of the normal mesh which consists of 692,064 elements. Furthermore,
Figure 2 shows the mesh model used in the current analysis.
3.4. Convergence Criteria
The numerical simulation was performed using the COMSOL Finite element software. At each iteration, the average relative error of the velocities, pressure, and temperature are each computed. These are obtained via the following relation:
Here, F represents one of the unknowns, viz., U, V, W, P, or θ, and s is the iteration number, and (i, j) represents the coordinates on the grid. Convergence is reached if R for all the unknowns is below 1 × 10−6 in two successive iterations.