The present study addresses the flow characteristics of a viscous, incompressible, steady, and Newtonian fluid flow through the undulating microchannel with a porous medium. The flow is governed by the Darcy–Brinkman model with no-slip boundary conditions at walls. The objective of this study is to develop theoretical and computational models for flow parameters that are independent of the permeability of the medium and to extend the scope of previous studies. The lubrication theory is used to determine key flow parameters, such as flow rate, velocity, and wall shear stress, in complex-shaped microchannels. To overcome the limitations of lubrication and boundary perturbation methods, the spectral method is applied to a sinusoidal microchannel. We observe that flow parameters are significantly affected by dimensionless quantities such as pattern amplitude, wavelength, and permeability ( ). The spectral model indicates non-linear flow rate behavior when the permeability is very high ( ) and accurately captures the transition behavior of flow rate in the Darcian flow regime for various wavelengths, unlike other theories. Conversely, for small and large wavelengths with low permeability ( ) at the Stokes flow limit, the flow rate behavior is monotonic. The spectral model demonstrates greater reliability compared to classical lubrication theory, extended lubrication theory, and boundary perturbation methods, especially for large values of the dependent variables. Predictions from the spectral approach closely align with numerical results over a broad range of parameters. A detailed analysis of the influence of various parameters on flow quantities is presented.
I. INTRODUCTION
The Darcy–Brinkmann (DB) model advances the understanding of fluid flow through a porous medium by extending the foundational principles of Darcy's law1 to include viscous shear effects. This enhancement is crucial for accurately analyzing flows that are significantly influenced by both the porous characteristics (Darcy effect1) and the viscous shear (Brinkmann effect2). Integration of these effects makes the DB model a more comprehensive tool for studying complex flow dynamics. Exploring fluid flow in porous media is essential for the design and optimization of various engineering systems, including filtration devices,3 chemical reactors, and technologies for cooling microelectronics.4,5 The DB model represents a foundational tool for understanding porous media flow. However, it is limited in addressing the complexities of flows through microchannels with patterned surfaces. In practical applications at the microscale regime, the surface is not smooth due to its porosity. Therefore, investigating the impact of surface patterning on DB flow is essential for understanding fluid behavior.
To explore the wide range of applications involving fluid flow in wavy channels with a fluid-saturated porous medium, efforts have been made to investigate the hydrodynamic phenomena and heat transfer processes in these systems.6–8 Additionally, flow separation phenomena have also been observed due to nonlinear geometries, as these shapes are crucial for mass transportation.9 Flow separation occurs normally when the local adverse pressure gradient is strong enough to alter the wall shear stress (WSS) or vorticity. Outside of fluid dynamics, pattern surfaces are also useful in heat transfer,10–12 microfluidic mixing,13–15 geoscience,16,17 etc. For example, the effect of pattern on forced convection was studied by Wang and Chen.18 In microscale, the roughness effect on heat transfer and pressure drop was investigated using the finite-element-based codes.19 The shape of the peaks was selected to control the surface roughness. Recently, the flow circulation near the crests has driven attention to flow through wavy channels, which significantly enhances fluid mixing.15,20–22
In the literature, the DB flow with micropatterning has been studied from different perspectives, such as different shapes of channel,23–27 no-slip,8,25,26 slip boundary conditions,28,29 Newtonian8,25,30 and non-Newtonian31,32 fluids, isotropic8,26,30 and anisotropic porosity25,33 of porous material, and flow direction. The investigation has been focused on the impact of wall corrugation on pressure drop, velocity distribution, and flow resistance in microchannels. Researchers have studied porous media flow in ducts and considered the DB model for the investigation, such as parallel plate channels,34 circular,35 rectangular ducts,23,24 and sinusoidal.8,30 The flow behavior has also been examined for two-dimensional8,30 and three-dimensional36,37 geometries. For all geometries, it is observed that initially, the velocity profile is flat, and it becomes parabolic when the porous media parameter is small. The DB equation was examined for a three-dimensional curved channel using the domain perturbation methods (DPMs), and it observed that the flow resistance strongly depends on the porous media factor and bumps of geometry.8 Furthermore, the flow of porous media was examined for a tube with a three-dimensional bumpy surface using the boundary perturbation method. It was assumed that the amplitude of the surface bumps is small compared to the mean tube radius.37 In this study, they found that flow resistance is minimized with permeability parameters and wave numbers. The aforementioned literature models assumed a no-slip boundary, and the permeability is considered to be isotropic porous media used to examine DB flow. These models are accurate up to small values of perturbation parameters (amplitude-to-pitch ratio). The porous media flow for two-dimensional channels with different orientations has also been studied by several researchers.8,26,30,38,39 The DB model is used to examine the non-Newtonian fluid flow through anisotropic porous media. Recently, Gupta and Vajravelu31 investigated non-Newtonian flow in an anisotropic rotating porous channel with an inclined magnetic field. They found that the maximum volumetric flow rate was obtained with higher values of the Jeffrey parameter. Bhargavi et al.32 studied the Darcy–Brinkman–Forchheimer model for a non-Newtonian fluid through anisotropic porous media. The study revealed that as the anisotropic parameter increases, the velocity profile transitions from a parabolic to a flatter shape at the center of the channel.
Several theoretical models have been proposed based on the flow direction8,25,30,38–40 such as longitudinal and transverse to study the porous media flow for different topographies. The eigenfunction expansion was used, and it was found that the flow resistance was lowest in the longitudinal flow.38 The boundary perturbation method was employed to examine DB flow through the grooved channel, and a mathematical model was developed for a flow rate for small pattern amplitude in longitudinal flow.8 In this study, they found that the effect of corrugation is greater in the Stokes limit than in the Darcian limit at a small wavelength. The flow orientation along the length has also been explored in the context of a rough, curved channel filled with a porous medium.26 This study examines the effects of channel curvature, wall corrugations, and medium permeability using the boundary perturbation method for small corrugation amplitudes. The porous medium significantly affects flow compared to a clear conduit, especially at low permeability. In porous media flow, the porosity of the channel plays a significant role in flow parameters. In the literature, the transverse flow8,37–39,41 has been given more attention than the longitudinal flow. The effect of the anisotropic permeability ratio for DB flow along the transverse direction has been examined, and the lubrication theory was employed to develop a theoretical model of an arbitrary channel.25 In this study, it was determined that the anisotropic permeability ratio causes flow separation near the wall crests, where viscous forces are significant.
A number of theoretical models were proposed using boundary perturbation method,8,26,30,38 lubrication theory,25,42,43 and Fourier methods44–49 to investigate the effect of micropatterning for Stokes flow and Darcy–Brinkmann flow. The aforementioned literature makes it clear that several studies have been conducted to understand the effect of micropatterning on DB flow. However, these studies are limited by small pattern amplitude, long channel wavelength, and a small permeability parameter. There is no study that examines this flow without restricting the dependent variables, such as permeability parameter, pattern amplitude, and wavelength. In the past, the domain perturbation method was applied to examine the effect of pattern surface on the Darcy–Brinkmann model, which is accurate at small values of perturbation parameter. The main purpose of the current work is to expand previous studies by Ng and Wang8 and Karmakar and Sekhar.25 Additionally, we found that the flow across the grooves has attracted considerably more attention than the flow along the grooves. To fill these gaps, we employed the lubrication theory for longitudinal flow to explore Darcy–Brinkmann flow for any arbitrary geometries. Later, this study is extended to a sinusoidal surface. Overall, the aim of the present research will be to develop mathematical and computational modeling for flow parameters, such as flow rate, velocity and wall shear stress, without the restriction of dependent variables. Therefore, a spectral approach is employed to conduct the present study. Additionally, a finite-element-based numerical simulation (NM) is also performed to understand the usefulness of the present models and the limitations of the existing models.
The present paper is structured as follows. In Sec. II, the key governing equations and boundary conditions are discussed. Mathematical modeling using lubrication and grid-free spectral approaches are thoroughly explained in Secs. III and IV, respectively. At the end of this section, an asymptotic model is calculated for a periodic sinusoidal surface. In Sec. V, we also briefly remark on special cases of spectral method such as Darcian flow limit, Stokes flow limit, and asymptotic solution. Section VI focuses on a relation of porous permeability between the lubrication and spectral model (SM). The behavior of flow rate, wall shear stress, and velocity profile with reference to dependent variables are represented graphically in Sec. VII. At the end of the paper, we discussed key conclusions, future work, and the application of the present work. Some key supporting equations and numerical simulations are discussed in Appendixes A and B, respectively.
II. THEORETICAL FORMULATION
The schematic illustrates a periodic grooved microchannel used for Darcy–Brinkman flow analysis. The flow direction is along the z axis (longitudinal flow). L, a, and h are defined as pattern wavelength, amplitude, and mean channel height, respectively. The top wall is considered flat for both approaches. (a) In the lubrication theory, the lower boundary is defined by an arbitrary shape, represented as . (b) In the spectral method, the lower boundary follows a sinusoidal profile, expressed as . In both approaches, the analysis assumes one periodic cell.
The schematic illustrates a periodic grooved microchannel used for Darcy–Brinkman flow analysis. The flow direction is along the z axis (longitudinal flow). L, a, and h are defined as pattern wavelength, amplitude, and mean channel height, respectively. The top wall is considered flat for both approaches. (a) In the lubrication theory, the lower boundary is defined by an arbitrary shape, represented as . (b) In the spectral method, the lower boundary follows a sinusoidal profile, expressed as . In both approaches, the analysis assumes one periodic cell.
III. LUBRICATION APPROXIMATION
A. Perturbation approximation and solution of the hydrodynamic problem
1. The leading order problem:
2. The second order problem:
B. Calculation for scaled flow rate
IV. SPECTRAL APPROACH
A. Spectral method
B. Calculation for scaled flow rate
V. NOVEL CONTRIBUTION OF THE SPECTRAL METHOD
A. Spectral equations are compared with the literature model: DB model
B. The spectral equations as compared with the previous domain perturbation models: Stokes flow
VI. RELATION BETWEEN CHARACTERIZING POROUS MEDIUM BY LUBRICATION THEORY AND SPECTRAL THEORY
VII. RESULTS AND DISCUSSION
The present study examines fluid flow through a channel with one grooved and one flat, impermeable wall, with porous material present in the channel. The fluid movement in the porous area is governed by the DB equation, considering only viscous forces. Lubrication equations are thoroughly validated by comparison with numerical simulations (NS) of the DB equation, as discussed in Sec. VII A. A finite-element-based numerical study is also conducted using COMSOL Multiphysics® software, as discussed in Appendix A. The present NS is validated with the existing literature models8,55 and the current theoretical models (lubrication approach and grid-free spectral approach). In this paper, numerical values are represented symbolically. This section effectively illustrates the influence of dependent variables on scaled flow rate through graphical representation.
The variation of the scaled flow rate ratio is shown against the non-dimensional pattern amplitude for different values of permeability parameters using CLT and ELT, as discussed in Sec. VII A. Additionally, the prediction from the domain perturbation method8 is cross-validated with the present NS, as discussed in Sec. VII B. The aforementioned sections reveal the limitations of existing models in the literature. Finally, Sec. VII C compares predictions from the spectral model (SM) with numerical simulations (NS) and discusses a non-linear behavior of scaled flow rate at different values of permeability parameters. The behavior of flow parameters (e.g., scaled flow rate, velocity, and wall shear stress) is examined with pattern amplitude, wavelength, and permeability of the porous medium.
A. Evaluating lubrication theory against numerical simulation
The impact of dimensionless pattern amplitude ( ) on the scaled flow rate for various values and perturbation parameters ( ) is depicted in Fig. 2. The solid and dashed curves illustrate the predictions from ELT and CLT, respectively. Equations (18) and (19) are utilized for CLT and ELT, respectively. Panel (a) of Fig. 2 demonstrates the variation of scaled flow rate with pattern amplitude ( ) for at . When the channel height (h) is 0.1 times channel length (L), the scaled flow rate ratio ( ) increases with pattern amplitude across a range of values. Both the numerical study and theoretical approaches (CLT and ELT) indicate an increasing trend in flow rate with , as depicted in panel (a) of Fig. 2. This comparison highlights the higher accuracy of ELT over CLT, and a close agreement is found between ELT and numerical values at both the Stokes flow limit ( ) and the Darcy flow limit ( ). CLT is accurate at smaller pattern amplitudes ( ). Conversely, when the channel height is 0.5 times a channel length ( ), the scaled flow rate decreases with increasing pattern amplitude, as illustrated in panel (b) of Fig. 2. This decreasing behavior is predicted by both ELT and NS. In contrast, CLT consistently predicts an increasing flow rate with pattern amplitude at both limits, which shows the unphysical behavior of the scaled flow rate. The ratio between h and L is second order in the governing equation [Eq. (3)], and the CLT provides the solution for , which is an incomplete investigation to capture the viscous stress. Consequently, the higher order correction in the CLT, which corresponds to a more accurate evaluation of viscous stresses, improves the applicability of lubrication theory from a significantly higher wavelength. Therefore, the higher-order term plays a significant role in the perturbation analysis to understand fluid behavior. Considering the higher-order term in the CLT accurately captures the viscous stress. Panel (b) indicates that the current theoretical models align closely with numerical values at smaller pattern amplitudes ( ) from the Stokes to Darcian flow limits.
Variation of the scaled flow rate with pattern amplitude across different permeability mediums ( ) is presented for and , as illustrated in panels (a) and (b), respectively. Equations (18) and (19) are used to show the prediction from the classical lubrication theory (CLT) and extended lubrication theory (ELT). Numerical simulations (NS) are represented symbolically.
Variation of the scaled flow rate with pattern amplitude across different permeability mediums ( ) is presented for and , as illustrated in panels (a) and (b), respectively. Equations (18) and (19) are used to show the prediction from the classical lubrication theory (CLT) and extended lubrication theory (ELT). Numerical simulations (NS) are represented symbolically.
The error percentage of the current theoretical model is assessed against numerical simulations to gauge accuracy. For example, at a perturbation parameter and a permeability parameter , ELT exhibits an error percentage of 2.644 , whereas CLT shows a significantly higher error percentage of 9.54 , both calculated at a pattern amplitude . When the perturbation parameter is increased to , ELT maintains an error percentage below 10 at and . However, the error percentage for ELT surpasses the 10 threshold at . It is observed that the relative difference between NS and ELT models increases as the pattern amplitude grows, indicating that the model's accuracy diminishes with larger pattern amplitudes. Section VII B presents a discussion of the results derived from the domain perturbation approach [ ], spectral method (SM), and numerical simulations (NS).
B. Comparative analysis of domain perturbation method and numerical simulation
Ng and Wang8 investigated Darcy–Brinkman flow through a sinusoidal grooved microchannel using the domain perturbation method (DPM). They developed a second-order solution for the scaled flow rate, providing a foundation for understanding the behavior of flow with reference to the pattern surface. This model is accurate at significantly small pattern amplitudes. The present study is an extension of the existing literature model and will develop a model for large pattern amplitudes. In this section, we compare different mathematical models, including DPM and SM, against finite-element-based numerical simulations (NS). Equation (31) represents the prediction from SM, whereas Eq. (34) is a prediction from DPM, which is accurate up to . By comparing SM and DPM with NS, we aim to evaluate the accuracy and consistency of these models against numerical results, providing deeper insights into their predictive capabilities.
Figure 3 illustrates the variation of the scaled flow rate ( ) as a function of pattern amplitude ( ). It compares predictions from the theoretical model with numerical results across various wavelengths ( ) and porous media permeabilities ( ). For a smaller wavelength ( ), an increase in leads to the reduced scaled flow rate in the Stokes flow limit , as shown in panel (a) of Fig. 3. This behavior is captured by all approaches. This compression suggests that the -solution is accurate at smaller values of pattern amplitude for , and it fails for large values of . In contrast, the prediction of scaled flow rate from SM is found to have a good agreement with numerical values at significantly large and . Furthermore, we analyzed for a large wavelength ( ), and the permeability of porous media varies from 0 to 1; increases with dimensionless pattern amplitude as shown in panel (c) of Fig. 3. At intermediate wavelength ( ) and , the flow rate ratio decreases monotonically with increasing . This trend is captured by only -solution, as shown in dotted lines in panel (b) of Fig. 3. By contrast, a non-monotonic behavior of is observed using SM [Eq. (31)] and numerical study at both values of . This observation suggests that the behavior of the scaled flow rate is non-linear, and the minimum flow rate is noticed at an intermediate wavelength for the Stokes and Darcian flow limits. For significantly large pattern amplitudes ( ), it is noticed that SM and NS exhibit a good agreement at both limiting flow, as indicated by the compression. The case of the large characterizing porous medium or Darcian flow limit is discussed in detail in Sec. VII C. We also discussed the increasing and decreasing behavior of the scaled flow rate in the later section.
Variation of scaled flow rate ( ) with pattern amplitude ( ) for different permeability mediums ( ) at different wavelengths ( ). Panels (a), (b), and (c) represent variations for , , and , respectively. Equations (31) and (34) illustrate the predictions derived from the spectral model (SM) and the asymptotic model [ ], respectively. Numerical simulations (NS) are denoted symbolically.
Variation of scaled flow rate ( ) with pattern amplitude ( ) for different permeability mediums ( ) at different wavelengths ( ). Panels (a), (b), and (c) represent variations for , , and , respectively. Equations (31) and (34) illustrate the predictions derived from the spectral model (SM) and the asymptotic model [ ], respectively. Numerical simulations (NS) are denoted symbolically.
C. Effect of permeability of porous media on scaled flow rate
This section examines the variation in the scaled flow rate with respect to pattern amplitude for large values of permeability parameter ( ), employing both SM and NS. The analysis is also conducted across various wavelengths ( ). Equation (31) is employed to plot the scaled flow rate as shown in panels (a) and (b) of Figs. 4 and 5. This compression clearly reveals that initially decreases with , followed by an increase as continues to rise at medium ( ) to large ( ) wavelengths. A similar trend is also captured by the numerical study. Figures 4 and 5 reveal that the parameter plays a crucial role in determining the transition behavior of the scaled flow rate. At the Stokes limit ( ) and , the non-monotonic behavior of the scaled flow rate is valid for intermediate wavelength ( ). By contrast, a similar trend is also noticed at Darcian flow at any values of wavelength ( ). As demonstrated in the literature model8 and the present lubrication theory model, the dimensional flow rate may exhibit either a monotonic increase or decrease with respect to the wavelength and permeability parameter, as discussed in Secs. VII A and VII B. However, in the spectral model's prediction, the dimensionless flow rate reveals non-linear behavior at both the Stokes and Darcy flow limits. The factors contributing to the observed fluctuations in the dimensional flow rate are examined at the end of this section. Figures 4 and 5 elucidated a good agreement between the numerical values and the prediction from SM. Panels (a) and (b) of Figs. 4 and 5 show that the SM prediction for the scaled flow rate always captured the minimum at large values of . Each wavelength with respect to has a different pattern amplitude, where flow rate behavior captured the minimum. The transition behavior of scaled permeability occurs at a critical pattern amplitude ( ). Table I shows the values of ; at this point, a non-linear behavior of scaled flow rate is observed. When , the flow rate through a grooved channel is lower than that of standard Poiseuille flow. Conversely, when , the flow rate through a grooved channel exceeds the standard Poiseuille flow rate. When is constant, for example, = 3, the critical pattern amplitude is ∼0.61 for = 0.5, which is lower than that observed for = 3.0. This finding indicates that increases with . A similar behavior is also captured for other values of .
Variation of scaled flow rate ( ) with pattern amplitude for different permeability ( ) mediums ( ) at different wavelengths ( ). Panels (a) and (b) are variations for and , respectively. The solid lines represent the prediction from the spectral model (SM), and Eq. (31) is used to obtain the variation of dimensional flow rate. Numerical simulations (NS) are represented symbolically.
Variation of scaled flow rate ( ) with pattern amplitude for different permeability ( ) mediums ( ) at different wavelengths ( ). Panels (a) and (b) are variations for and , respectively. The solid lines represent the prediction from the spectral model (SM), and Eq. (31) is used to obtain the variation of dimensional flow rate. Numerical simulations (NS) are represented symbolically.
Variation of dimensional flow rate ( ) with pattern amplitude for different permeability ( ) mediums ( ) at different wavelengths ( ). Panels (a) and (b) are variations for and , respectively. The solid lines represent the prediction from the spectral model (SM), and Eq. (31) is used to obtain the variation of the scaled flow rate. Numerical simulations (NS) are represented symbolically.
Variation of dimensional flow rate ( ) with pattern amplitude for different permeability ( ) mediums ( ) at different wavelengths ( ). Panels (a) and (b) are variations for and , respectively. The solid lines represent the prediction from the spectral model (SM), and Eq. (31) is used to obtain the variation of the scaled flow rate. Numerical simulations (NS) are represented symbolically.
The transition behavior of dimensional flow rate ( ) is calculated with critical pattern amplitude ( ) at various wavelengths ( ) and permeability mediums ( ). The boldface value denotes approximate values of critical pattern amplitude, where the dimensionless flow rate is calculated and transmission behaviour is captured.
. | . | . | . |
---|---|---|---|
3 | 0.5 | 0.61 | 0.9494 |
3 | 3 | 0.80 | 0.9298 |
10 | 15 | 0.58 | 0.9974 |
10 | 25 | 0.79 | 0.9967 |
. | . | . | . |
---|---|---|---|
3 | 0.5 | 0.61 | 0.9494 |
3 | 3 | 0.80 | 0.9298 |
10 | 15 | 0.58 | 0.9974 |
10 | 25 | 0.79 | 0.9967 |
We can draw the following conclusions based on the trends observed in Figs. 4 and 5, as well as the aforementioned discussion regarding the impact of , , and . As increases, there is an increase in the additional area on the lower wall [as shown in panel (b) of Fig. 1] over which viscous forces can enforce the no-slip condition. This increase in area is higher than that of a channel with a flat lower wall at Y = 0. When the wavelength of the pattern is short, the most significant gradients of the flow occur near the lower wall and are inversely proportional to the wavelength. This amplifies the role of the viscous forces acting on the lower surface for the DB flow. Due to this effect, the scaled flow rate decreases with at short wavelengths. For short waves, the grooves behave as roughness elements that enhance drag. However, in the long-wave limit, the x-gradients in the flow, which are influenced by the wavelength, become less significant than the y-gradients that are determined by the local channel thickness. As a result, when is small initially and at moderate to large wavelengths with significantly large , grooves behave as roughness elements that enhance friction. This leads to a negative slope in the flow rate vs amplitude curve, as seen in Figs. 4 and 5. On the other hand, when is increased, a transition is observed in the flow rate vs amplitude curve. Initially, a minimum is reached, as seen in the figures. However, on further increasing , the slope of the curve changes toward a long wave-like positive slope. This is because the y-gradients become more significant than x-gradients due to the formation of progressively narrower constrictions with an increase in .
1. Wall shear stress (WSS) distribution near to the grooved wall
This study aims to elucidate the effects of pattern amplitude ( ) and permeability parameter ( ) on wall shear stress (WSS) behavior along a grooved wall. We present a graphical representation of WSS distribution in Figs. 6–8. The investigation is performed for small ( ) and long ( ) wavelengths of the channel. Panels (a) and (b) of Fig. 6 illustrate the variation of WSS across different values of at a fixed value of . If the channel length (L) equals the mean channel height (h) or , WSS reaches its maximum value at the center of the channel ( ) for all values of . The maximum value of WSS increases with increasing pattern amplitude, as illustrated in panel (a) of Fig. 6. The viscous forces are more prominent near the crest of the grooved surface for a shorter wavelength. For the case where ( ) with , the maximum wall WSS is observed at for and , as shown in panel (b) of Fig. 6. However, for a slightly larger pattern amplitude ( ), the maximum WSS decreases, and its location shifts away from . This indicates that the viscous effects become less significant as the channel length increases relative to its height.
Variation of wall shear wall stress is calculated on the bottom grooved wall for = 15 and various values of . Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically.
Variation of wall shear wall stress is calculated on the bottom grooved wall for = 15 and various values of . Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically.
Variation of wall shear wall stress is calculated on the bottom grooved wall for = 0.4 and various values of . Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically, whereas the solid lines are prediction from the spectral method (SM).
Variation of wall shear wall stress is calculated on the bottom grooved wall for = 0.4 and various values of . Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically, whereas the solid lines are prediction from the spectral method (SM).
Variation of wall shear wall stress is calculated on the bottom grooved wall for = 0.4 and various values of . Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically, whereas the solid lines are prediction from the spectral method (SM).
Variation of wall shear wall stress is calculated on the bottom grooved wall for = 0.4 and various values of . Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically, whereas the solid lines are prediction from the spectral method (SM).
Figure 7 graphically demonstrates the variation of WSS for different values of (0.1, 1.0, 5.0, 10.0, 15.0, and 20.0) at for smaller and large wavelength. If and , the maximum WSS is observed in the Stokes limit ( ), while the minimum WSS is identified in the Darcian limit ( ), as illustrated in panel (a) of Fig. 7. Conversely, for , the maximum WSS is found in the Darcian limit, and the minimum WSS is observed in the Stokes limit, as shown in panel (b) of Fig. 7. The viscous shear effect is highly pronounced in the Stokes flow regime and diminishes with increasing permeability of the porous material. Consequently, the maximum wall shear stress (WSS) is observed in the Stokes flow limit compared to the Darcian flow limit, particularly at small wavelengths. This observation indicates that the maximum and minimum WSS values occur at the narrowest section of the channel, corresponding to . Furthermore, we observed that an increase in the permeability parameter significantly affects intermediate wavelengths, such as 3 and 5. Consequently, the influence of WSS on the grooved wall surface becomes less prominent. The distribution of WSS is similar for and as shown in panels (a) and (b) of Fig. 8. When the channel has a more permeable and shorter wavelength, the viscous effect is less significant. This means there is greater resistance on the grooved wall. As a result, WSS decreases with a more permeable medium. On the other hand, viscous forces are more significant for long wavelengths and larger permeable surfaces. These findings show that amplitude and wavelength strongly depend on the geometrical properties of a porous material. The prediction from the spectral method is perfectly aligned with the numerical values for each case.
2. Variation of scaled flow rate with reference to permeability of porous media ( )
This section examines the influence of on the variation of scaled flow rate. To elucidate its significance, we utilize both spectral method and numerical simulation approaches. Figure 9 demonstrates the transition from the Stokes flow to the Darcian flow. Equation (31) presents the prediction of the scaled flow rate using the spectral approach, while Eq. (34) provides the solution for the scaled flow rate. The comparison between the analytical solution and numerical study for = 3 and = 5 is depicted in panels (a) and (b), respectively. At the Stokes and Darcy flow limits, the results from the current spectral method reveal a good agreement with numerical simulations across all pattern amplitude values. At , the scaled flow rate first decreases and then increases as increases, as shown in panel (a). When varies from 0 to 10 with , the behavior of the scaled flow rate exhibits a decreasing trend, followed by an increasing trend, which can be attributed to the effects of surface roughness and confinement, respectively. Both the numerical simulation and the solution show a similar trend; however, a clear difference exists between them. In contrast, for , the scaled flow rate steadily decreases as increases. At moderately large wavelengths, viscous forces become more pronounced with increasing values of . The permeability of the porous medium reduces the scaled flow rate. In the Stokes-to-Darcian flow regime ( ), the flow rate through the grooved channel (Q) matches that of a straight channel ( ). This behavior is observed from all approaches, as shown in panel (b). As seen in Fig. 9, larger channel wavelengths are associated with higher flow rates in both Stokes and Darcy flow limits. In contrast, moderate wavelengths show a combination of increasing and decreasing flow rates in these two regimes. A non-linear phenomenon is observed at 1.885. For , the observed decreasing trend is attributed to the predominant influence of surface roughness. In contrast, for , the increasing trend indicates the confinement effects are becoming more significant. When analyzing DB flow in a grooved channel, it is essential to consider the channel's wavelength. This factor affects flow behavior and its interaction with the patterned surface, influencing parameters such as pressure drop, flow distribution, and overall channel performance.
The effect of permeability medium for the dimensionless flow rate ( ) is shown by the spectral method, numerical simulation, and asymptotic results of Ng and Wang.8 Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically, whereas the solid lines stand for the prediction from the spectral method (SM).
The effect of permeability medium for the dimensionless flow rate ( ) is shown by the spectral method, numerical simulation, and asymptotic results of Ng and Wang.8 Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically, whereas the solid lines stand for the prediction from the spectral method (SM).
3. Variation of scaled flow rate with reference to the height of the channel
The variation of the scaled flow rate is plotted with scaled minimum spacing between grooved and straight plates for a small amplitude-to-pitch ratio ( ) for various permeability medium as shown in Fig. 10. The logarithmic scale is used for the independent variable to capture the low-channel height regime and the minimum in each curve precisely. Similarly, a large amplitude-to-pitch ( ) ratio is also investigated here as depicted in Fig. 11. When the flat wall (top wall) is brought closer from a large distance to a patterned surface, two competing phenomena are observed with various amplitude-to-pitch ratios. For example, when the permeability is small ( ), the dimensional flow rate is larger for a grooved channel as compared to a straight channel. A similar trend is also observed with different values of permeability as shown in Fig. 10. The flow rate ratio decreases with increasing values of permeability if it is scaled with rather than h. The flow rate ratio is always higher than the standard Poiseuille flow through a straight channel. For a large amplitude-to-pitch, the flow rate ratio is always less than that of the standard Poiseuille flow through a straight channel. In this situation, the non-linear behavior of dimensionless flow rate is identified with various permeability values, as illustrated in Fig. 11. This transition behavior of flow rate can be summarized dimensionally. When the dimensional amplitude-to-pitch ratio ( ) is 3 and L = 10 μm, then amplitude (a) is equal to 4.75 μm with different permeability ( ) values such as 0.05, 0.5, 1, and 2. Using these quantities, the scaled flow rate reaches a minimum of 45.67 , 56.65 , 69.29 , and 81.88 of the flow rate of the plane channel flow with the same h and different , when h is 30.6 μm. Smaller and larger channel sizes than h = 30.6 μm will result in a larger flow rate ratio to the plane channel flow with the same h. The aforementioned discussion shows that the flow rate strongly depends on the permeability of porous media and the dimension of the pattern amplitude.
At the Stokes to Darcian flow limit, variation of dimensionless flow rate ( ) with a scaled minimum spacing between patterned and unpatterned plates for the amplitude-to-pitch ratio ( ). A logarithmic scale is used for the independent variable to visually resolve the low-channel height regime and the minimum in each curve. Equation (31) is used to obtain the variation of the scaled flow rate.
At the Stokes to Darcian flow limit, variation of dimensionless flow rate ( ) with a scaled minimum spacing between patterned and unpatterned plates for the amplitude-to-pitch ratio ( ). A logarithmic scale is used for the independent variable to visually resolve the low-channel height regime and the minimum in each curve. Equation (31) is used to obtain the variation of the scaled flow rate.
At the Stokes to Darcian flow limit, variation of dimensionless flow rate ( ) with a scaled minimum spacing between patterned and unpatterned plates for the amplitude-to-pitch ratio ( ). Equation (31) is used to obtain the variation of the scaled flow rate.
At the Stokes to Darcian flow limit, variation of dimensionless flow rate ( ) with a scaled minimum spacing between patterned and unpatterned plates for the amplitude-to-pitch ratio ( ). Equation (31) is used to obtain the variation of the scaled flow rate.
4. Variation of scaled flow rate with wavelengths
The plot in Fig. 12 depicts how the scaled flow rate changes with wavelengths ( ) for a fixed value of 0.9. This situation is highly constrained, indicating that the thinnest section of the channel is 19 times bigger than the narrowest portion. An investigation is conducted into the Stokes flow limit, as illustrated in panel (a), and the Darcian flow limit, as depicted in panel (b). Equation (31) is used for the prediction by SM, whereas Eq. (34) is considered for an -solution. Both theories are compared with NS. A good agreement is found between SM and NS for each value of wavelengths at both limits. For , the flow rate is lower than that of the conventional Poiseuille flow at smaller wavelengths, while it exceeds the Poiseuille flow at larger wavelengths as shown in panel (a). A smaller wavelength is emphasized by the roughness effect, while a longer wavelength is associated with the confinement effect. For , the flow rate is always higher than that of the conventional Poiseuille flow at any values of wavelength as shown in panel (b). The model is not accurate at smaller to large wavelengths, whereas the prediction from Eq. (31) is accurate at any value of wavelength for Stokes limit and Darcian limit, as shown in both panels of Fig. 12.
The effect of wavelength for the dimensionless flow rate ( ) is shown by the spectral method, numerical simulation, and asymptotic results of Ng and Wang.8 Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically, whereas the solid lines stand for the prediction from the spectral method (SM).
The effect of wavelength for the dimensionless flow rate ( ) is shown by the spectral method, numerical simulation, and asymptotic results of Ng and Wang.8 Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically, whereas the solid lines stand for the prediction from the spectral method (SM).
5. Variation of longitudinal velocity
To gain a better understanding of the role played by in the current problem, it is essential to analyze the axial velocity profile. Equation (22) is not only helpful to identify the flow rate through a channel, but it is also utilized to understand other flow behavior such as flow profile (velocity). The depth-wise variation of velocity at different transverse locations within the flow cross section is plotted in Figs. 13 and 14. For Darcian flow, the analysis is performed for both small ( ) and large ( ) wavelengths. The results focus on the thickest part of the channel ( ), as illustrated in panels (a) and (b) of Fig. 13. However, Fig. 14 presents the flow profile at the thinnest section of the channel ( ) for various values of . At various values of , the comparison in these figures clearly shows that the theoretical distribution matches the numerical distribution in the Y direction when 0.5. At , the velocity profile deviates from a parabolic shape, exhibiting a flattened characteristic in the center of the channel across all values of , as depicted in panel (a) of Fig. 13. It is observed that the axial velocity decreases with increasing . Higher values of reduce the effective permeability in the flow direction, thereby increasing the overall resistance to flow. As a result, the fluid velocity diminishes with increasing . Furthermore, an increase from 2 to 10 results in a reduction of velocity at the thickest section of the channel. With a wavelength of = 10 and = 6, the depth-wise velocity distribution displays a parabolic profile. The maximum velocity occurs at the center of the channel, as illustrated in panel (b) of Fig. 13. When increases from 6 to 15, the velocity profile deviates from the parabolic shape. For larger values of , the axial velocity profile becomes increasingly uniform, with the influence of the wall confined to a thin boundary layer near the surface. As a result, the wall resistance has a minimal contribution to the overall resistance, which is predominantly governed by the porous medium. This results in a decrease in the overall velocity. This comparison signifies the dominance of Darcian flow over the Stokes flow at different wavelengths. To conclude, if the length of the channel (L) is 10 times the mean height (h), it is essential to employ reduced values of in order to achieve a parabolic velocity profile.
At various permeability mediums, the variation of velocity profile along the transverse direction is shown at the thickest section of that channel ( ) for moderately large pattern amplitude ( ). Equation (22) is used to obtain the variation of the scaled flow rate. Panel (a) corresponds to , while panel (b) corresponds to . Panels (c) and (d) illustrate the variation of axial velocity at and for , = 2, 4. The solid lines represent the predictions obtained using the spectral method (SM), while the symbols correspond to the results from the numerical simulations (NS).
At various permeability mediums, the variation of velocity profile along the transverse direction is shown at the thickest section of that channel ( ) for moderately large pattern amplitude ( ). Equation (22) is used to obtain the variation of the scaled flow rate. Panel (a) corresponds to , while panel (b) corresponds to . Panels (c) and (d) illustrate the variation of axial velocity at and for , = 2, 4. The solid lines represent the predictions obtained using the spectral method (SM), while the symbols correspond to the results from the numerical simulations (NS).
At various permeability mediums ( ), the variation of velocity profile along the transverse direction is shown at the thinnest section of that channel ( ) for moderately large pattern amplitude ( ). Equation (22) is used to obtain the variation of the scaled flow rate. Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically.
At various permeability mediums ( ), the variation of velocity profile along the transverse direction is shown at the thinnest section of that channel ( ) for moderately large pattern amplitude ( ). Equation (22) is used to obtain the variation of the scaled flow rate. Panel (a) corresponds to , while panel (b) corresponds to . Numerical simulations (NS) are represented symbolically.
Panels (c) and (d) illustrate the variation in the axial velocity for two values of (2 and 4) at two different channel sections ( and ) for and . The results indicate that as increases, the velocity magnitude decreases, particularly at the thickest section of the channel ( ). In regions with high material porosity, the flow velocity is reduced, and the velocity profile transitions from parabolic to flatter, as shown in panel (c) of Fig. 13. In contrast, at the thinnest section of the channel, the velocity profile remains close to parabolic, and the maximum velocity W values are nearly identical for = 2 and 4, as shown in panel (d). Overall, has a more pronounced effect at shorter wavelengths compared to longer wavelengths. A strong agreement is observed between the spectral and numerical approaches.
At the thinnest section of the channel ( ), the velocity profile is not parabolic with reference to any values of at = 1, as illustrated in panel (a) of Fig. 14. For a large wavelength, the velocity profile is always captured in a parabolic nature or similar to Poiseuille flow as illustrated in panel (b) of Fig. 14. The influence of becomes more pronounced when the length of the channel (L) is equal to its mean height (h). Conversely, when the channel length is ten times the mean height, the effect of diminishes significantly. Overall, the magnitude of the velocity (W) decreases with increasing at any wavelength of the channel. The aforementioned discussion reveals that the flow profiles strongly depend on and . A good agreement is found between the prediction from the spectral model and the numerical simulation. Furthermore, the no-slip boundary condition at both walls is achieved using spectral method and numerical simulations.
The analysis focuses on the behavior of the axial velocity (W) at the Stokes flow limit ( ) for both the thickest ( ) and thinnest ( ) sections of the channel, considering both small ( ) and large wavelengths ( ), as illustrated in Fig. 15. The investigation is conducted for various values of , including 0.1, 0.3, 0.5, and 1. For and each value of , the maximum velocity is observed at the center of the channel at both sections, as shown in panels (a) and (c). At , the velocity profile closely resembles standard Poiseuille flow (parabolic profile). As increases from 0.1 to 0.5 and then from 0.5 to 1, the fluid velocity in the porous medium decreases, with the velocity profile becoming progressively flatter at the center of the channel, approaching the Darcian flow limit. Panels (a) and (c) of Fig. 15 indicate that flow resistance increases with permeability at both channel sections, and the influence of the porous material becomes more pronounced slightly away from the wall. For and , similar trends are observed, as shown in panel (b) of Fig. 15. However, at large wavelengths, the effect of is less pronounced compared to the small wavelength case. The minimum axial velocity is observed at = 1. At , the effect of the permeability of the porous media is less significant compared to , as shown in panel (d). The maximum velocity for the transition from Stokes to Darcian flow is nearly identical for both sections. For all cases, excellent agreement is found between the spectral method and numerical simulations, as shown in Fig. 15. The above-mentioned analysis demonstrates that the presence of a porous medium effectively suppresses the growth of boundary layers on rigid, impermeable plates, preventing them from extending to the center of the channel. This behavior contrasts sharply with that of plane Poiseuille flow, where boundary layers develop uniformly from the plates and merge at the channel center, as indicated by . Consequently, the introduction of a porous medium distinctly modifies the flow dynamics compared to conventional cases.
Variation of axial velocity in the Stokes flow regime is examined at different locations within the channel, specifically at the thinnest ( ) and thickest ( ) sections for and . The solid lines represent predictions from the special method (SM), while the symbols denote results from the numerical simulations (NS).
Variation of axial velocity in the Stokes flow regime is examined at different locations within the channel, specifically at the thinnest ( ) and thickest ( ) sections for and . The solid lines represent predictions from the special method (SM), while the symbols denote results from the numerical simulations (NS).
Figure 16 presents both the contour and surface plots showing the variation in the magnitude of the velocity (W) with different values of for and . The figure indicates that as increases from the Stokes to Darcian flow limit, the fluid velocity decreases, particularly at the thickest section of the channel ( ), due to the influence of permeability in the porous medium. For instance, at the Stokes flow limit = 0, the maximum velocity is ∼0.224, while for = 5, it decreases to around 0.055. In contrast, at the thinnest part of the channel ( ), the maximum velocity is nearly the same (0.003) for both the Stokes and Darcy limits. This observation suggests that the permeability effect is more pronounced in the thicker sections of the channel than in the thinner sections, particularly for long wavelengths. Additionally, the maximum velocity is observed at the Stokes flow limit, as compared to the Darcy flow limit. In high-permeability conditions, flow resistance in wave channels becomes more pronounced, counteracting the applied pressure gradient. As a result, the velocity profile tends to flatten.
For a large wavelength ( ) and pattern amplitude ( ), panels (a)–(h) illustrate the variation in velocity magnitude (W) for different under the Stokes and Darcian flow limits.
For a large wavelength ( ) and pattern amplitude ( ), panels (a)–(h) illustrate the variation in velocity magnitude (W) for different under the Stokes and Darcian flow limits.
VIII. CONCLUSION
In this paper, we present a comprehensive study of hydrodynamic flow through a corrugated channel containing sparse porous material. The primary objective is to investigate Darcy–Brinkman flow in a grooved microchannel, focusing on cases with large pattern amplitudes and high permeability in the porous medium, particularly for complex geometries. The analysis is conducted using both spectral and lubrication approaches: the lubrication theory is applied to complex topographies, while spectral methods are utilized for sinusoidal channels. While previous models in the literature, based on the domain perturbation theory, offer accurate predictions for small pattern amplitudes and low permeability, their applicability diminishes as these parameters increase. Furthermore, studies on Darcy–Brinkman flow have predominantly focused on transverse flow, with limited attention given to longitudinal flow in porous media. To address these challenges, we propose a mathematical model that ensures an accurate scaled flow rate even in scenarios with high permeability and large pattern amplitudes. This study employs a systematic perturbation analysis to derive higher-order terms, extending the traditional lubrication approximation. The analytical results are thoroughly compared with numerical simulations, demonstrating the accuracy and reliability of the proposed model.
The key conclusions are as follows: (i) In both the Stokes and Darcy limits, the higher-order extended lubrication theory (ELT) shows excellent agreement with numerical simulations at , demonstrating a significant improvement over the classical lubrication theory (CLT). However, in the case of , ELT fails with increasing pattern amplitude. (ii) The spectral model overcomes these limitations of the lubrication approximation at both limiting flows. Based on the Fourier theory, the current proposed model is more accurate than the domain perturbation theory, regardless of dependent parameters. We developed a model that is accurate when the crest of a grooved wall is very close to the upper wall. (iii) According to the literature model, the dimensional flow rate may increase or decrease with dependent parameters, which provides an incomplete behavior for the Darcy–Brinkmann flow through pattern topographies. The present spectral model predicts a non-monotonic behavior of the scaled flow rate at the Darcian limit and provides a broader view of the effect of micropatterning on Darcy–Brinkmann flow. The non-linear behavior of the scaled flow rate is a significant finding of the present work, whereas the literature models are unable to capture this trend. The Darcy viscous drag force is very prominent and depends on the geometry's orientation. (iv) At smaller wavelengths, the maximum shear stress is observed at the crest of the pattern surface for Stokes flow ( ) and Darcian flow limit ( ). In contrast, for a large wavelength, the wall shear stress behavior significantly alters with the permeability medium. The amplitude-to-pitch ratio significantly influences the range of maximum and minimum shear stress values. (v) At the Stokes flow limit, the velocity profile has a parabolic shape, similar to standard Poiseuille flow. However, in the case of a high permeability medium ( ) or the Darcian limit, the profile becomes flatter.
Equations (22) and (31) are key equations to predict the behavior of flow parameters for Darcy–Brinkman flow through a grooved microchannel. We have demonstrated the influence of dependent variables on the scaled flow rate with micropatterning. The findings of the present study are highly significant and have broad applications in multiple fields, such as chemical reactors, filtering equipment, heat exchangers, and flow through rock fractures.
ACKNOWLEDGMENTS
The author would like to thank Dr. Tim Persoons, Department of Mechanical, Manufacturing and Biomedical Engineering, Trinity College Dublin, Dublin, Ireland, for providing access to the software.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Mainendra Kumar Dewangan: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Investigation (lead); Methodology (lead); Validation (lead); Writing – original draft (lead); Writing – review & editing (lead).
DATA AVAILABILITY
The data that support the findings of this study are available within the article.
APPENDIX A: NUMERICAL SIMULATION
Finite-element analysis using COMSOL Multiphysics is employed to evaluate the present model and assess the limitations of existing literature models. The partial differential equation (PDE) module in COMSOL is utilized to solve the governing equations and boundary conditions. The PARDISO (parallel direct sparse solver) direct solver, a stationary solver with relative tolerance was used to obtain the numerical values. For the analysis, a single periodic cell of the microchannel is considered, with periodic boundary conditions applied along the y axis within the computational domain. Numerical simulation is performed for the lubrication approach and spectral approach separately. For the lubrication theory, Eq. (3) with Eqs. (4a) and (4b) is solved using the PDEs module in the software, whereas, for the spectral method, Eq. (20) with Eqs. (21a) and (21b) is considered for the solution. No slip boundary conditions at walls (X-edges) and periodic boundary conditions at Y-edges are chosen. The domain is discretized using a triangular mesh, with mesh refinement ensuring accuracy up to 3–4 decimal places to evaluate mesh dependency. A grid sensitivity analysis is performed to examine how variations in the mesh size affect the computed flow quantities. This includes an independent grid study for each wavelength and pattern amplitude. The size of the computational cells depends on the maximum and minimum element sizes, element size growth rate, and maximum allowed iterations. We found that further refining the mesh did not affect the accuracy of the numerical study. The maximum and minimum element sizes are chosen based on the grid independence test as depicted in Table II. In the numerical simulations (NS), we have assessed the limitations of the computational framework. Our findings indicate that the NS are sufficiently robust to handle extreme scenarios, such as (i) cases where the crest of the grooved surface approaches the upper flat surface ( ) and (ii) situations where the permeability of the porous medium is exceptionally high ( ). The spectral model is shown to provide faster predictions compared to the numerical simulations, highlighting its computational efficiency.
Variation of scaled flow rate at , , and for different mesh sizes. The boldface value denotes that the grid sensitivity is accurate up to 4 decimal points.
Element size . | Number of cell count . | . |
---|---|---|
0.2 | 791 | 0.991 65 |
0.1 | 1 084 | 0.995 92 |
0.08 | 1 637 | 0.996 47 |
0.05 | 4 115 | 0.996 87 |
0.03 | 11 314 | 0.996 94 |
0.02 | 25 255 | 0.996 95 |
0.01 | 101 011 | 0.996 96 |
Element size . | Number of cell count . | . |
---|---|---|
0.2 | 791 | 0.991 65 |
0.1 | 1 084 | 0.995 92 |
0.08 | 1 637 | 0.996 47 |
0.05 | 4 115 | 0.996 87 |
0.03 | 11 314 | 0.996 94 |
0.02 | 25 255 | 0.996 95 |
0.01 | 101 011 | 0.996 96 |