Next Article in Journal
Design and Operational Assessment of a Railroad Track Robot for Railcar Undercarriage Condition Inspection
Next Article in Special Issue
Optimizing Selective Laser Melting of Inconel 625 Superalloy through Statistical Analysis of Surface and Volumetric Defects
Previous Article in Journal
Mapping the Potential of Zero-Energy Building in Greece Using Roof Photovoltaics
Previous Article in Special Issue
Investigation on the Microstructure and Micro-Mechanical Properties of Thermal-Sprayed NiCoCrAlY High Entropy Alloy Coating
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Computational Investigation of the Fluidic Properties of Triply Periodic Minimal Surface (TPMS) Structures in Tissue Engineering

by
Muhammad Noman Shahid
1,
Muhammad Usman Shahid
1,
Shummaila Rasheed
1,
Muhammad Irfan
1 and
Muhannad Ahmed Obeidi
2,*
1
Department of Mechanical Engineering, Capital University of Science and Technology, Islamabad 45750, Pakistan
2
School of Mechanical and Manufacturing Engineering, Dublin City University, D09 V209 Dublin, Ireland
*
Author to whom correspondence should be addressed.
Designs 2024, 8(4), 69; https://doi.org/10.3390/designs8040069
Submission received: 16 June 2024 / Revised: 7 July 2024 / Accepted: 8 July 2024 / Published: 10 July 2024
(This article belongs to the Special Issue Post-manufacturing Testing and Characterization of Materials)

Abstract

:
Tissue engineering, a rapidly advancing field in medicine, has made significant strides with the development of artificial tissue substitutes to meet the growing need for organ transplants. Three-dimensional (3D) porous scaffolds are widely utilized in tissue engineering, especially in orthopedic surgery. This study investigated the fluidic properties of diamond and gyroid structures with varying porosity levels (50–80%) using Computational Fluid Dynamics (CFD) analysis. The pressure and velocity distributions were analyzed, and it was observed that the pressure decreased gradually, whereas the velocity increased in the central area of the surface structures. Specifically, the pressure drop ranged from 2.079 to 0.984 Pa for the diamond structure and from 1.669 to 0.943 Pa for the gyroid structure as the porosity increased from 50% to 80%. It was also found that the permeability increased as the porosity level increased, with values ranging from 2.424 × 10 9 to 5.122 × 10 9   m 2 for the diamond structure and from 2.966 × 10 9 to 5.344 × 10 9   m 2 for the gyroid structure. The wall shear stress (WSS) was also analyzed, showing a consistent decrease with increased porosity for both types of structures, with WSS values ranging from 9.903 × 10 2 to 9.840 × 10 1   P a for the diamond structure and from 1.150 × 10 1 to 7.717 × 10 2   P a for the gyroid structure. Overall, this study provides insights into the fluidic properties of diamond and gyroid structures, which can be useful in various applications such as tissue engineering.

1. Introduction

In recent years, tissue engineering has seen substantial advancements, driven by the creation of novel artificial tissue replacements designed to address the increasing demand for organ transplants. Particularly in orthopedic surgery, three-dimensional (3D) porous scaffolds with diverse structural designs are frequently employed in tissue engineering. Advancements in manufacturing technologies, including additive manufacturing, have enabled the production of 3D scaffolds with a variety of materials and structures. Nevertheless, achieving the ideal scaffold design remains challenging, as the design parameters significantly influence the scaffold’s mechanical properties and biocompatibility [1,2]. For example, while a high porosity is favorable for biological compatibility, it may compromise the mechanical integrity of the scaffold. It is also necessary to properly design the mass flow characteristics of porous structures, such as permeability, to permit the passage of nutrients and oxygen into cells inside the structure [3,4].
In tissue engineering, it is a common practice to predict the permeability of scaffolds using computer models. Studies have shown that porous structures can serve as scaffolds to facilitate osteogenic differentiation, adhesion, proliferation, host integration, cell attachment, and cell viability. To facilitate nutrient diffusion, cell infiltration, and the removal of metabolic waste products, scaffolds must exhibit certain characteristics. They should have interconnected pores and a high level of porosity [5]. It is difficult to develop a porous structure that meets the requirements of both mechanical and biological functions. However, the geometry and curvature of the surfaces where cells reside are recognized as crucial factors in evaluating the rate of tissue production. Recent theories suggest that, unlike flat and convex surfaces, high-curvature surfaces and concave surfaces, respectively, promote and enhance tissue regeneration [6]. Additionally, surface roughness is an important quality measure as it supports cell differentiation and growth, while also potentially localizing stress. These requirements can be addressed by designing scaffolds based on Triply Periodic Minimal Surfaces (TPMS) [3,7].
A TPMS, or the surface of the minimal area, possesses a boundary delineated by a closed curve [8]. TPMS scaffolds are constructed on this surface, with each unit cell created by adding thickness to the surface and subsequently combining to form a final scaffold that exhibits cubic symmetry and comprises periodically repeated interconnected pores. Therefore, a TPMS is more desirable for bone regeneration [9], and has characteristics similar to the curvature of the trabecular bone [10]. Several natural geometries such as cubosomes, cell membranes, and plant prolamellar bodies can be characterized by a TPMS [7]. Previous research has investigated how various pore sizes impact the bond strength between an implant and bone. Their findings indicated that pore sizes between 300 μm and 600 μm were optimal in facilitating continuous bone growth to enhance osseointegration. The largest pore size demonstrated the quickest fluid movement, indicating that increasing pore size can enhance the rate of cell migration [11].
Evaluating the characteristics related to biocompatibility, such as the permeability and effect of wall shear stress, is of the utmost importance because of their impact on the biological activity of cells within the supportive structures [12]. Numerous investigations have been undertaken to measure the permeability of scaffolds and analyze the factors that impact this permeability. A scaffold that allows substances such as nutrients and gasses to pass through its pores adequately while also enabling the effective removal of waste is considered permeable [13]. Pires et al. found that calculating permeability is crucial for optimizing cell cultivation and determining treatment duration, as it directly influences cell growth and proliferation within the supportive structures [14]. Ma et al. revealed that, by modifying the design parameters, it is possible to adjust the permeability and mechanical characteristics of the Gyroid structure to mimic those found in trabecular bones within the human body. Furthermore, they observed that permeability had a significant impact on initial cell growth [15]. When designing a scaffold, permeability is a vital consideration, as it significantly influences the transport of oxygen, nutrients, and the removal of metabolic waste. Geometric factors, including pore size, surface thickness, and structure type, also affect permeability [5,16,17]. Researchers have investigated the concept of developing the microstructure of scaffolds through topology optimization. The findings demonstrated a direct relationship between computational and experimental permeability, underscoring the significance of permeability as a crucial design factor. The study illustrated that the regeneration of tissue is influenced by factors such as porosity and permeability, highlighting their potential for furthering our understanding of the impact of scaffold architecture on biological processes [18]. Aqila et al. [19] conducted a study to examine how the size of pores influences the mechanical behavior and fluid permeability of bone scaffolds. Their findings indicated that larger pore sizes were associated with improved permeability and reduced wall shear stress across different flow rates, facilitating a more effective transportation of fluids and nutrients. Ntousi et al. [20] used a computational fluid dynamic (CFD) analysis to analyze two scaffold structures with a pore size of 500 μm. The results revealed that the designed scaffolds exhibited enhanced permeability and fluid flow, making them advantageous for bone tissue engineering applications. Pires et al. [14] evaluate the permeability of various TPMS scaffold geometries through a combination of experimental pressure drop tests and CFD simulations. The results suggested that Gyroid structures were the most optimal choice for applications in bone tissue engineering, as they consistently demonstrated fluid permeation, closely followed by Schwarz D. Conversely, Schwarz P configurations displayed uniform flow patterns and notable fluctuations in permeability based on porosity levels.
Another critical factor in studying the fluidic response of scaffolds is Wall Shear Stress (WSS). WSS has been recognized for its impact on cell growth [8,21]. Fluid-induced WSS provides a mechanical stimulus that can be sensed by cells, leading to enhanced cell differentiation and proliferation [22]. Ali et al. [23] conducted a comparative analysis using CFD simulations to assess the permeability and WSS of TPMS and lattice-based scaffolds. The analysis revealed how internal structures affect the behavior of fluids. Lesman et al. [24] conducted CFD simulations to demonstrate that when cells grow within 3D scaffolds during perfusion cell culturing, the WSS increases as a result of contraction in the channels of the scaffolds. Marin et al. [25] conducted a CFD analysis to investigate how the variation in structure among scaffolds affects the distribution and magnitude of WSS. The findings of their research revealed that the presence of geometric inconsistencies led to significant variations in velocity and WSS across the different samples. A CFD study was conducted in scaffolds with square pores of varying sizes, with a constant porosity of 63% to evaluate the permeability and WSS. The research revealed that s higher surface roughness led to a decline in both permeability and WSS. It is noteworthy that there was a notable decrease in WSS observed when comparing smooth and rough scaffolds [3]. Zhao et al. [26] conducted a study that analyzed the impact of the geometric characteristics of scaffolds using a fluid–structure interaction model. The results showed that pore size significantly influences mechanical stimulation, and simultaneous fluid perfusion and mechanical load increase WSS. In addition, Lee et al. [27] utilized the Allen–Cahn equation to observe and analyze the dynamic growth behaviors of cells within various scaffold structures. Although various authors have explored the impacts of the geometric aspects of TPMS structures on mechanical performance [22,28,29], studying the fluid behavior of different TPMS structures remains a promising area of research, particularly in applications related to bone implants.
The primary objective of this research is to create TPMS structures utilizing nTopology 4.24.3 software and subsequently optimize them. Following this, the optimized scaffolds, namely Diamond and Gyroid, which have different levels of porosity, are evaluated to determine their fluidic flow characteristics. CFD analysis is employed to examine physical characteristics including pressure differentials, permeability, and WSS. These parameters provide a basis for investigating the biological behavior of these structures. Our investigation focuses on evaluating the flow characteristics of these structures. The goal is to assess their potential as bone scaffolds and contribute to the advancement of optimized bone tissue engineering applications.

2. Materials and Methods

Initially, a scaffold was developed leading to the creation of a three-dimensional model. Following these procedural steps, a CFD analysis was undertaken on the scaffolds, investigating the fluidic behavior of blood under varying porosity levels.

2.1. Scaffold Design

The design of scaffolds is critical in CFD studies because of the behavior of fluidic flow, which is dependent on lattice structure. Two architectural scaffolds, namely diamond and gyroid, were selected and developed utilizing the nTopology software having different porosities ranging from 50% to 80% with an increment of 10% (Figure 1). The scaffolds were subsequently exported to ANSYS SpaceClaim, where they were converted from STL to solids. The different porosities in the diamond and gyroid structures are achieved by altering the wall thickness shown in Table 1 Thicker walls reduce the void space, resulting in lower porosity, while thinner walls increase the void space, leading to higher porosity. The cells display symmetry along each of the three orthogonal axes (x, y, and z), making them isotropic in their geometric properties.
CFD analysis was conducted using the ANSYS Fluent 2021 R2 module. To streamline computational time, each model was maintained at a size of 2 mm × 2 mm × 2 mm, taking into account the geometric complexity of the model. The 2 mm × 2 mm × 2 mm size is optimal for computational models due to its balance between efficiency and capturing the geometric intricacies of bone scaffolds [6]. While not reflecting all clinical scenarios, it represents a subset for analyzing fundamental fluid behavior within computational constraints. For larger scaffold structures, we emphasize the importance of cautious evaluation and experimental validation. Our study aims to provide critical insights on fundamental fluid dynamics principles, which are essential for a wide range of clinical scaffold sizes. In Figure 1, the subscript indicates the porosity percentage, like D 50 which represents 50% porosity.

2.2. Governing and Adopted CFD Equations

The following approach utilized the Navier–Stokes equation, taking into account an incompressible fluid with constant density.
ρ u t μ 2 u + ρ u . u + p = F
. u = 0
where ρ , u, μ, ∇, p, and F represent the density of the fluid ( k g m 3 ) , the fluid velocity (m/s), dynamic viscosity of fluid (kg/m.s), the del operator, pressure (Pa), and other forces (gravity or centrifugal force; in this case F = 0), respectively [6]. The blood density was assumed to remain constant at 1060 k g / m 3 for the CFD models.
The Carreau–Yasuda model was used to compute the fluid viscosity for the analysis, considering blood as a non-Newtonian fluid. The Carreau–Yasuda model was chosen for its ability to describe viscosity plateaus in bone scaffold applications, capturing fluid behavior under varying porosities. It proved essential in predicting pressure drop-flow rate relationships, demonstrating effectiveness in analyzing Carreau fluid flow through porous media [30]. The model’s accuracy in predicting fluid behavior makes it a valuable tool for understanding and optimizing blood flow through intricate scaffold structures, aiding in the design of efficient bone tissue engineering systems.
μ = μ + μ 0 μ [ 1 + ( λ γ ˙ ) 2 ] ( n 1 ) / 2
The equation utilized for Carreau fluids generalized from Darcy law is presented below:
Δ p L = U η e f f k = U k ( μ + μ 0 μ ) [ 1 + ( λ γ ˙ ) 2 ] ( n 1 ) / 2
where u and L denote the inlet fluid velocity (m/s) and model length (m), respectively. The values of μ and μ 0 represent the shear rate limit viscosities when the shear rate is either infinite or zero, respectively, relaxation time constant is λ, and the power law index is n [30].
In a laminar flow system, the WSS, denoted as τω, can be mathematically expressed as the normal velocity gradient on a wall.
τ ω = μ u m
where m indicates the x, y, and z directions [16].
The Reynolds number for non-Newtonian fluid can be calculated by [31]:
R e = ρ u L μ a
where ρ , u , L , and μ a denote density, velocity of fluid, length, and viscosity of fluid.
The Reynolds number for this non-Newtonian fluid under the given conditions is approximately 0.0052, which indicates a laminar flow regime. The characterization of blood flow as laminar in the CFD analysis of TPMS structures is based on physiological considerations, primarily the low flow velocities within blood vessels, resulting in Reynolds numbers conducive to laminar flow. The relatively high viscosity of blood and the small-scale nature of the vessels further support the assumption of laminar flow. This modeling choice aligns with typical flow conditions in the human circulatory system, enhancing the credibility of the simulation results.

2.3. Boundary Conditions

The study employed a laminar flow of fluid with an inlet velocity of 0.0001 m/s [6,17]. The direction of flow was chosen based on its potential to simplify the computational setup and boundary conditions, leading to a more accurate analysis and understanding of outcomes. Figure 2 shows the boundary conditions. To specify different fluid properties at the surface, such as velocity and pressure, a wall boundary condition was employed. In this study, a no-slip condition was imposed on the porous structure, which was assumed to be hydrophilic, consistent with typical practices in bone tissue engineering applications [25,32]. This assumption accounts for wetting behavior, leading to more accurate and practical CFD simulations. Furthermore, the outlet pressure was set to zero [15], to simplify the pressure differential computation, which was a critical parameter.

2.4. Numerical Schemes

In this study, the Finite Volume Method (FVM) has been employed as the numerical framework to simulate blood flow through the scaffold structures. To account for the non-Newtonian behavior of blood, the Carreau–Yasuda model has been integrated, offering a comprehensive description of viscosity variations with shear rate. The governing equations were solved using a coupled scheme, which simultaneously solves the momentum and continuity equations to enhance convergence stability and efficiency, particularly beneficial for flow interactions within the complex geometries of the scaffolds. For spatial discretization, the Green Gauss Node Based method was employed for gradient calculations due to its superior accuracy in handling unstructured meshes. The pressure interpolation was conducted using a second-order scheme to ensure a higher accuracy in capturing pressure variations within the flow domain. Momentum equations were discretized using a first-order upwind scheme, which, despite being less accurate than higher-order schemes, provides robust and stable solutions, particularly in the initial stages of simulation or in regions with high gradients and complex flow patterns. This combination of numerical schemes was chosen to balance computational efficiency and solution accuracy, ensuring reliable simulation results for the analysis of blood flow behavior through the TPMS scaffolds.

2.5. Grid Convergence

The study focused on the characterization of blood as a non-Newtonian fluid. A mesh sensitivity analysis was conducted to assess how variations in mesh element sizes influenced the results. Figure 3 shows a visual representation of the mesh in the fluid domain containing the scaffold. Mesh elements with sizes of 0.2, 0.5, and 1 mm were used to compare the permeability and WSS values of a specific model. The findings revealed that using a mesh size of 0.2 mm resulted in a minimal variation, less than 2%, in the WSS of the unit cell. As a result, a size of element of 0.2 mm was selected for the ongoing investigation. Furthermore, calculations were performed for WSS, permeability, pressure, and fluid velocity. The spatial discretization is accomplished using a tetrahedral structured mesh. To ensure the model’s performance sensitivity, results were brought into convergence by establishing a residual criterion value of 1 × 10−6.

3. Numerical Model Validation

The present numerical methodology is validated by comparing the results with the literature. In this regard, the permeability results for the Gyroid structure are compared with the Ali and Sen [16] study. They analyzed the Gyroid structure using the power law model with an inlet velocity of 0.7 mm/s. The results from the present methodology match well with those from the Ali and Sen [16] study with a reasonable accuracy. Figure 4 shows the validation of our numerical methodology for the permeability of a Gyroid structure.

4. Results and Discussions

4.1. Fluidic Characteristics of Scaffolds

An analysis of pressure and velocity distribution was conducted to examine the fluidic characteristics of Diamond (D) and Gyroid (G) structures with varying porosity levels ranging from 50% to 80%. The pressure distribution contours of the D surface (i.e., D 50 , D 60 , D 70 , and D 80 ) and the G surface (i.e., G 50 , G 60 , G 70 , and G 80 ) are presented in Figure 5. It is evident that there is a gradual decrease in pressure from the inlet to the outlet in all structures, regardless of the porosity level. The observed pressure variation in TPMS-based scaffolds, with the higher pressure near the inlet diminishing as the fluid progresses downwards, is a consequence of the intricate interplay between geometric complexities and fluid dynamics. The initial compression near the inlet, induced by the complex scaffold geometry, leads to regions of elevated pressure. Subsequent fluid flow through smoother paths and porous structures results in a gradual pressure dissipation. At the bottom of each picture, negative values of the pressure appear. This is because, near the outlet, the speed was increasing (Figure 5) so the pressure is decreased. Also, due to the effect of gravity, the speed increases as the fluid progresses so the pressure decreased.
The velocity distribution within scaffolds, as illustrated in Figure 6, reveals the profound impact of internal design on flow characteristics. It is crucial to observe that the central regions of scaffold structure surface have velocities that exceed the initial inlet velocity. The increase in speed can be attributed to the complex lattice geometry, which causes a localized rise in fluid velocity. At the same time, adjacent regions demonstrate the combined influence of geometric features and fluid dynamics [15]. The observed acceleration of the fluid profile, particularly in the central areas, has significant implications for the transportation of nutrients to deeper structures within the lattice. The improved fluid dynamics in these central regions suggest a potential optimization for the dispersion of nutrients, which agrees with results of the previous studies [33,34].
Figure 7 shows the pressure drop versus porosity, as by increasing the porosity, the pressure drop decreases in both the Diamond and Gyroid structures. The Diamond structure shows a higher pressure drop than the Gyroid structure due to its higher stiffness and strength, which result in a lower permeability. A study investigated the mechanical properties of TPMS-based Diamond and Gyroid Ti6Al4V scaffolds for bone implants. The findings revealed that, when comparing scaffolds with identical pore sizes, the Diamond scaffold exhibited a greater stiffness and strength than the Gyroid scaffold. As a result, the Diamond scaffold exhibited a higher pressure drop [35]. This phenomenon can be attributed to the inherent differences in the geometrical characteristics and flow dynamics of the two structures. The complex geometry of the Diamond structure induces more resistance to fluid flow, resulting in a higher pressure drop, whereas the Gyroid structure, with its distinct architecture, presents a comparatively lower resistance, leading to a reduced pressure drop. Figure 7 shows that the pressure drop ranges from 2.079 Pa to 0.984 Pa for the Diamond structure and 1.669 Pa to 0.943 Pa for the Gyroid structure for porosities ranging from 50% to 80%.
The TPMS structures are made from stainless steel, which has a high Young’s modulus (200 GPa) and yield strength (500 MPa), making them rigid and resistant to deformation under fluid loads. Additionally, the small dimensions (a few millimeters) and thick walls of the structures ensure structural integrity, minimizing deformation. Similar studies, such as those by Song et al. [36], have also assumed non-deformable vessels for TPMS structures due to their rigidity and structural design, supporting our assumption.

4.2. Permeability

Equation (4) was used to calculate the permeability of the Diamond and Gyroid structures, considering the pressure difference obtained through the CFD analysis. To assess the transport efficiency of different porous structures, both permeability (k) and pressure difference (∆p) were taken into account.
For the successful occurrence of mass transfer within any porous structure, permeability is of the utmost importance. In Figure 8, the expected permeability of both the Diamond and Gyroid surfaces is depicted for the porosity range of 50% to 80%. It is noted that a higher porosity results in an increased permeability because the pores in the scaffold allow for better fluid flow. Research has indicated that TPMS scaffolds with higher porosity levels have higher permeability. The Gyroid scaffold type has been determined to be more permeable than the Diamond scaffold type [37]. The continuous, three-dimensional architecture of the Gyroid offers a larger surface area for the flow of fluids and facilitates a more interconnected pathway for fluid passage. These characteristics of the Gyroid, including its enhanced interconnectivity and greater void space, contribute to its higher permeability, allowing fluids to move through the Gyroid scaffold with reduced resistance compared to the Diamond scaffold. Ali et al. [23] studied Gyroid- and lattice-based architectures with porosities ranging from 65% to 90% to observe their permeability. Our results show a good agreement with the results from this study. Gómez S. et al. [38] observed the properties of 3D scaffolds for bone tissue engineering. They showed that the properties of scaffolds can be controlled in design stage to match exactly with trabecular bone. Van Bael S. [39] performed a study to observe the effect of pore size and permeability on Ti6Al4V scaffolds of three different pore shapes at 500 μm and 1000 μm. Nauman E. et al. [40] conducted a study to observe the permeability of trabecular bone for tissue engineering applications. Beaudoin et al. [41] observed the permeability of cancellous bone.
In this study, the permeability ranged from 2.424 × 10 9   m 2 to 5.122 × 10 9   m 2 for the Diamond structure and 2.966 × 10 9   m 2 to 5.344 × 10 9   m 2 for the Gyroid structure. Table 2 shows the good agreement of the obtained results with previously reported results.

4.3. WSS

Equation (5) was utilized to determine the WSS of diamond and gyroid structures. The analysis of the WSS on the Gyroid and Diamond structures revealed a consistent trend. WSS consistently decreased with an increase in porosity for both types of structures, which attributed to the reduced surface area available for fluid interaction within the porous framework.
When comparing the WSS values between the two TPMS structures, the Diamond TPMS exhibited consistently lower WSS values compared to the Gyroid TPMS across various porosities shown in Figure 9. The results show that the Diamond TPMS is more effective in reducing WSS compared to the Gyroid TPMS, which is beneficial for improving the overall performance of the TPMS structure. This difference in WSS values between the two TPMS structures may have significant implications for their performance in different applications. Figure 9 shows that the WSS ranges from 9.903 × 10 2     P a to 9.840 × 10 1     P a for the Diamond structure and 1.150 × 10 1   P a to 7.717 × 10 2   P a for the Gyroid structure.

5. Conclusions

In this study, the fluidic properties of Diamond and Gyroid TPMS-based scaffolds with varying porosities (50–80%) were examined through CFD analysis, focusing on their potential applications in orthopedic implants. The findings highlight significant differences in pressure distribution, permeability, and WSS between the two structures. Gyroid scaffolds consistently demonstrated superior fluidic performance compared to Diamond structures, with lower pressure drops and higher permeability across all porosity levels. These characteristics suggest that Gyroid scaffolds are more conducive to nutrient flow and waste removal, and their gradual decrease in WSS with increased porosity indicates a favorable environment for cell proliferation and tissue regeneration. Therefore, Gyroid structures are recommended as the preferred choice for orthopedic implants applications due to their balanced fluidic properties and potential to support biological processes effectively.
Although the results were derived from a model with limited dimensions (2 mm × 2 mm × 2 mm), they provide foundational insights that could be applicable to larger scaffold samples. Future research should focus on scaling these models and validating them with experimental data to ensure their applicability to clinically relevant scaffold sizes. The necessity of experimental validation is evident; while the simulations offer a comprehensive understanding of fluidic behavior, physical experiments, such as flow perfusion bioreactor studies, are required to confirm the predicted pressure drops, permeability, and WSS values under physiological conditions. Additionally, experimental studies could explore the biological responses, such as cell adhesion and growth, within these scaffolds under dynamic flow conditions.

Author Contributions

Conceptualization, M.N.S., M.U.S., S.R., M.I. and M.A.O.; formal analysis, M.N.S., M.U.S., S.R., M.I. and M.A.O.; funding acquisition, S.R. and M.A.O.; investigation, M.N.S., M.U.S., S.R. and M.I.; methodology, M.N.S., M.U.S. and S.R.; project administration, S.R.; supervision, S.R. and M.I.; writing—original draft, M.N.S. and M.U.S.; writing—review and editing, S.R. and M.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yadav, P.; Beniwal, G.; Saxena, K.K. A Review on Pore and Porosity in Tissue Engineering. Mater. Today Proc. 2021, 44, 2623–2628. [Google Scholar] [CrossRef]
  2. Rasheed, S.; Lughmani, W.A.; Obeidi, M.A.; Brabazon, D.; Ahad, I.U. Additive Manufacturing of Bone Scaffolds Using Polyjet and Stereolithography Techniques. Appl. Sci. 2021, 11, 7336. [Google Scholar] [CrossRef]
  3. Ali, D.; Sen, S. Computational Fluid Dynamics Study of the Effects of Surface Roughness on Permeability and Fluid Flow-Induced Wall Shear Stress in Scaffolds. Ann. Biomed. Eng. 2018, 46, 2023–2035. [Google Scholar] [CrossRef]
  4. Rasheed, S.; Lughmani, W.A.; Khan, M.M.; Brabazon, D.; Obeidi, M.A.; Ahad, I.U. The Porosity Design and Deformation Behavior Analysis of Additively Manufactured Bone Scaffolds through Finite Element Modelling and Mechanical Property Investigations. J. Funct. Biomater. 2023, 14, 496. [Google Scholar] [CrossRef]
  5. Castro, A.P.G.; Pires, T.; Santos, J.E.; Gouveia, B.P.; Fernandes, P.R. Permeability versus Design in TPMS Scaffolds. Materials 2019, 12, 1313. [Google Scholar] [CrossRef]
  6. Kumar, J.; Nirala, N.S.; Singh, N.K.; Gupta, N.; Dwivedi, Y.D.; Verma, R.; Rai, S.K.; Gupta, M. Design, Development and Fluidic Behavior Analysis of Triply Periodic Minimal Surface (TPMS) Based Scaffolds for Bone-Applications. Int. J. Interact. Des. Manuf. 2023, 1–11. [Google Scholar] [CrossRef]
  7. Wang, Z.; Huang, C.; Wang, J.; Wang, P.; Bi, S.; Abbas, C.A. Design and Simulation of Flow Field for Bone Tissue Engineering Scaffold Based on Triply Periodic Minimal Surface. Chin. J. Mech. Eng. 2019, 32, 19. [Google Scholar] [CrossRef]
  8. Poltue, T.; Karuna, C.; Khrueaduangkham, S.; Seehanam, S.; Promoppatum, P. Design Exploration of 3D-Printed Triply Periodic Minimal Surface Scaffolds for Bone Implants. Int. J. Mech. Sci. 2021, 211, 106762. [Google Scholar] [CrossRef]
  9. Abbasi, N.; Hamlet, S.; Love, R.M.; Nguyen, N.T. Porous Scaffolds for Bone Regeneration. J. Sci. Adv. Mater. Devices 2020, 5, 1–9. [Google Scholar] [CrossRef]
  10. Restrepo, S.; Ocampo, S.; Ramirez, J.A.; Paucar, C.; Garcia, C. Mechanical Properties of Ceramic Structures Based on Triply Periodic Minimal Surface (TPMS) Processed by 3D Printing. J. Phys. Conf. Ser. 2017, 935, 012036. [Google Scholar] [CrossRef]
  11. Barba, D.; Alabort, E.; Reed, R.C. Synthetic Bone: Design by Additive Manufacturing. Acta Biomater. 2019, 97, 637–656. [Google Scholar] [CrossRef] [PubMed]
  12. Rahbari, A.; Montazerian, H.; Davoodi, E.; Homayoonfar, S. Predicting Permeability of Regular Tissue Engineering Scaffolds: Scaling Analysis of Pore Architecture, Scaffold Length, and Fluid Flow Rate Effects. Comput. Methods Biomech. Biomed. Engin 2017, 20, 231–241. [Google Scholar] [CrossRef] [PubMed]
  13. Chabanon, M.; Duval, H.; Grenier, J.; Beauchesne, C.; Goyeau, B.; David, B. Histological Method to Study the Effect of Shear Stress on Cell Proliferation and Tissue Morphology in a Bioreactor. Tissue Eng. Regen. Med. 2019, 16, 225–235. [Google Scholar] [CrossRef] [PubMed]
  14. Pires, T.; Santos, J.; Ruben, R.B.; Gouveia, B.P.; Castro, A.P.G.; Fernandes, P.R. Numerical-Experimental Analysis of the Permeability-Porosity Relationship in Triply Periodic Minimal Surfaces Scaffolds. J. Biomech. 2021, 117, 110263. [Google Scholar] [CrossRef]
  15. Ma, S.; Tang, Q.; Han, X.; Feng, Q.; Song, J.; Setchi, R.; Liu, Y.; Liu, Y.; Goulas, A.; Engstrøm, D.S.; et al. Manufacturability, Mechanical Properties, Mass-Transport Properties and Biocompatibility of Triply Periodic Minimal Surface (TPMS) Porous Scaffolds Fabricated by Selective Laser Melting. Mater. Des. 2020, 195, 109034. [Google Scholar] [CrossRef]
  16. Ali, D.; Sen, S. Permeability and Fluid Flow-Induced Wall Shear Stress of Bone Tissue Scaffolds: Computational Fluid Dynamic Analysis Using Newtonian and Non-Newtonian Blood Flow Models. Comput. Biol. Med. 2018, 99, 201–208. [Google Scholar] [CrossRef] [PubMed]
  17. Ali, D.; Ozalp, M.; Blanquer, S.B.G.; Onel, S.; Permeability, S.O. Permeability and Fluid Flow-Induced Wall Shear Stress in Bone Scaffolds with TPMS and Lattice Architectures: A CFD Analysis. Eur. J. Mech. B/Fluids 2020, 79, 376–385. [Google Scholar] [CrossRef]
  18. Hollister, S.J. Porous Scaffold Design for Tissue Engineering. Nat. Mater. 2005, 4, 518–524. [Google Scholar] [CrossRef] [PubMed]
  19. Aqila, N.; Hussein, K.; Aziah, N.; Mohtar, J.; Haqim Neza, L.; Noordin, M.A.; Putra, A.; Saad, M. Pore Size Effect on Mechanical Response and Fluid Permeability for Bone Scaffold Regeneration. Mal. J. Med. Health Sci. 2021, 17, 34–39. [Google Scholar]
  20. Ntousi, O.; Roumpi, M.; Siogkas, P.; Deligianni, D.; Fotiadis, D.I. Computational Fluid Dynamic Analysis of Customised 3D-Printed Bone Scaffolds with Different Architectures. In Proceedings of the 2023 45th Annual International Conference of the IEEE Engineering in Medicine & Biology Society (EMBC), Sydney, Australia, 24–27 July 2023; Volume 2023. [Google Scholar] [CrossRef]
  21. Guyot, Y.; Papantoniou, I.; Luyten, F.P.; Geris, L. Coupling Curvature-Dependent and Shear Stress-Stimulated Neotissue Growth in Dynamic Bioreactor Cultures: A 3D Computational Model of a Complete Scaffold. Biomech. Model. Mechanobiol. 2016, 15, 169–180. [Google Scholar] [CrossRef]
  22. Maskery, I.; Sturm, L.; Aremu, A.O.; Panesar, A.; Williams, C.B.; Tuck, C.J.; Wildman, R.D.; Ashcroft, I.A.; Hague, R.J.M. Insights into the Mechanical Properties of Several Triply Periodic Minimal Surface Lattice Structures Made by Polymer Additive Manufacturing. Polymer 2018, 152, 62–71. [Google Scholar] [CrossRef]
  23. Ali, D.; Sen, S. Finite Element Analysis of Mechanical Behavior, Permeability and Fluid Induced Wall Shear Stress of High Porosity Scaffolds with Gyroid and Lattice-Based Architectures. J. Mech. Behav. Biomed. Mater. 2017, 75, 262–270. [Google Scholar] [CrossRef] [PubMed]
  24. Lesman, A.; Blinder, Y.; Levenberg, S. Modeling of Flow-Induced Shear Stress Applied on 3D Cellular Scaffolds: Implications for Vascular Tissue Engineering. Biotechnol. Bioeng. 2010, 105, 645–654. [Google Scholar] [CrossRef]
  25. Marin, A.C.; Lacroix, D. The Inter-Sample Structural Variability of Regular Tissue-Engineered Scaffolds Significantly Affects the Micromechanical Local Cell Environment. Interface Focus 2015, 5, 20140097. [Google Scholar] [CrossRef] [PubMed]
  26. Zhao, F.; Vaughan, T.J.; McNamara, L.M. Quantification of Fluid Shear Stress in Bone Tissue Engineering Scaffolds with Spherical and Cubical Pore Architectures. Biomech. Model. Mechanobiol. 2016, 15, 561–577. [Google Scholar] [CrossRef]
  27. Lee, H.G.; Park, J.; Yoon, S.; Lee, C.; Kim, J. Mathematical Model and Numerical Simulation for Tissue Growth on Bioscaffolds. Appl. Sci. 2019, 9, 4058. [Google Scholar] [CrossRef]
  28. Lu, Y.; Zhao, W.; Cui, Z.; Zhu, H.; Wu, C. The Anisotropic Elastic Behavior of the Widely-Used Triply-Periodic Minimal Surface Based Scaffolds. J. Mech. Behav. Biomed. Mater. 2019, 99, 56–65. [Google Scholar] [CrossRef] [PubMed]
  29. Sewify, G.H.; Javid, K.; Adeel, M.; Abbasi, A.; Khan, S.U.; Omri, M.; Kolsi, L. Blood Flow in Multi-Sinusoidal Curved Passages with Biomimetic Rheology: An Application of Blood Pumping. Mathematics 2022, 10, 1579. [Google Scholar] [CrossRef]
  30. Shahsavari, S.; McKinley, G.H. Mobility of Power-Law and Carreau Fluids through Fibrous Media. Phys. Rev. E Stat. Nonlin Soft Matter Phys. 2015, 92, 063012. [Google Scholar] [CrossRef]
  31. Cho, Y. Effects of the Non-Newtonian Viscosity of Blood on Hemodynamics of Diseased Arterial Flows. Biorheology 1991, 28, 241–262. [Google Scholar] [CrossRef]
  32. Truscello, S.; Kerckhofs, G.; Van Bael, S.; Pyka, G.; Schrooten, J.; Van Oosterwyck, H. Prediction of Permeability of Regular Scaffolds for Skeletal Tissue Engineering: A Combined Computational and Experimental Study. Acta Biomater. 2012, 8, 1648–1658. [Google Scholar] [CrossRef] [PubMed]
  33. Ma, S.; Tang, Q.; Feng, Q.; Song, J.; Han, X.; Guo, F. Mechanical Behaviours and Mass Transport Properties of Bone-Mimicking Scaffolds Consisted of Gyroid Structures Manufactured Using Selective Laser Melting. J. Mech. Behav. Biomed. Mater. 2019, 93, 158–169. [Google Scholar] [CrossRef] [PubMed]
  34. Jain, T.; Jain, J.K.; Saxena, K.K. Design and Comprehensive Study of Biodegradable Zinc-Based Implants for Bio-Medical Applications. Adv. Mater. Process. Technol. 2022, 8, 519–536. [Google Scholar] [CrossRef]
  35. Naghavi, S.A.; Tamaddon, M.; Marghoub, A.; Wang, K.; Babamiri, B.B.; Hazeli, K.; Xu, W.; Lu, X.; Sun, C.; Wang, L.; et al. Mechanical Characterisation and Numerical Modelling of TPMS-Based Gyroid and Diamond Ti6Al4V Scaffolds for Bone Implants: An Integrated Approach for Translational Consideration. Bioengineering 2022, 9, 504. [Google Scholar] [CrossRef] [PubMed]
  36. Song, J.; Wang, M.; Li, D.; Zhang, J. Deformation and Energy Absorption Performance of Functionally Graded TPMS Structures Fabricated by Selective Laser Melting. Appl. Sci. 2024, 14, 2064. [Google Scholar] [CrossRef]
  37. Santos, J.; Pires, T.; Gouveia, B.P.; Castro, A.P.G.; Fernandes, P.R. On the Permeability of TPMS Scaffolds. J. Mech. Behav. Biomed. Mater. 2020, 110, 103932. [Google Scholar] [CrossRef] [PubMed]
  38. Gómez, S.; Vlad, M.D.; López, J.; Fernández, E. Design and Properties of 3D Scaffolds for Bone Tissue Engineering. Acta Biomater. 2016, 42, 341–350. [Google Scholar] [CrossRef] [PubMed]
  39. Van Bael, S.; Chai, Y.C.; Truscello, S.; Moesen, M.; Kerckhofs, G.; Van Oosterwyck, H.; Kruth, J.P.; Schrooten, J. The Effect of Pore Geometry on the in Vitro Biological Behavior of Human Periosteum-Derived Cells Seeded on Selective Laser-Melted Ti6Al4V Bone Scaffolds. Acta Biomater. 2012, 8, 2824–2834. [Google Scholar] [CrossRef]
  40. Nauman, E.A.; Fong, K.E.; Keaveny, T.M. Dependence of Intertrabecular Permeability on Flow Direction and Anatomic Site. Ann. Biomed. Eng. 1999, 27, 517–524. [Google Scholar] [CrossRef]
  41. Beaudoin, A.J.; Mihalko, W.M.; Krause, W.R. Finite Element Modelling of Polymethylmethacrylate Flow through Cancellous Bone. J. Biomech. 1991, 24, 127–136. [Google Scholar] [CrossRef]
Figure 1. TPMS-based scaffold models with porosities of 50–80%: (ad) Diamond; (eh) Gyroid.
Figure 1. TPMS-based scaffold models with porosities of 50–80%: (ad) Diamond; (eh) Gyroid.
Designs 08 00069 g001aDesigns 08 00069 g001b
Figure 2. Boundary conditions for CFD analysis.
Figure 2. Boundary conditions for CFD analysis.
Designs 08 00069 g002
Figure 3. Visual representation of mesh in fluid domain containing the scaffold.
Figure 3. Visual representation of mesh in fluid domain containing the scaffold.
Designs 08 00069 g003
Figure 4. Validation of numerical methodology for the permeability of Gyroid structure [16].
Figure 4. Validation of numerical methodology for the permeability of Gyroid structure [16].
Designs 08 00069 g004
Figure 5. Pressure distribution contours of diamond structure (left side) and gyroid structures (right side).
Figure 5. Pressure distribution contours of diamond structure (left side) and gyroid structures (right side).
Designs 08 00069 g005
Figure 6. Velocity streamlines the distribution of diamond structures (left side) and gyroid structures (right side).
Figure 6. Velocity streamlines the distribution of diamond structures (left side) and gyroid structures (right side).
Designs 08 00069 g006aDesigns 08 00069 g006b
Figure 7. Pressure drop vs. porosity.
Figure 7. Pressure drop vs. porosity.
Designs 08 00069 g007
Figure 8. Permeability vs. porosity.
Figure 8. Permeability vs. porosity.
Designs 08 00069 g008
Figure 9. WSS vs. porosity.
Figure 9. WSS vs. porosity.
Designs 08 00069 g009
Table 1. The wall thickness of TPMS-based scaffolds.
Table 1. The wall thickness of TPMS-based scaffolds.
TPMS Based Scaffolds50%60%70%80%
Diamond wall thickness (mm)0.420.330.250.16
Gyroid wall thickness (mm)0.510.410.310.21
Table 2. Permeability of porous scaffolds and bone.
Table 2. Permeability of porous scaffolds and bone.
Ali D. et al. [23]Gómez S. et al. [38]Van Bael S. et al. [39]Nauman E. A. et al. [40]Beaudoin A. J. et al. [41]Current Work
Permeability   ( 10 9   m 2 ) 1.0 36.0   # 5.0 45.0   # 5.06 30.50   # 0.0268 20   * 0.467 14.80   * 2.424–5.122
P o r o u s   s c a f f o l d   # ;   B o n e   * .
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shahid, M.N.; Shahid, M.U.; Rasheed, S.; Irfan, M.; Obeidi, M.A. Computational Investigation of the Fluidic Properties of Triply Periodic Minimal Surface (TPMS) Structures in Tissue Engineering. Designs 2024, 8, 69. https://doi.org/10.3390/designs8040069

AMA Style

Shahid MN, Shahid MU, Rasheed S, Irfan M, Obeidi MA. Computational Investigation of the Fluidic Properties of Triply Periodic Minimal Surface (TPMS) Structures in Tissue Engineering. Designs. 2024; 8(4):69. https://doi.org/10.3390/designs8040069

Chicago/Turabian Style

Shahid, Muhammad Noman, Muhammad Usman Shahid, Shummaila Rasheed, Muhammad Irfan, and Muhannad Ahmed Obeidi. 2024. "Computational Investigation of the Fluidic Properties of Triply Periodic Minimal Surface (TPMS) Structures in Tissue Engineering" Designs 8, no. 4: 69. https://doi.org/10.3390/designs8040069

Article Metrics

Back to TopTop