Next Article in Journal
An Insight into Blood Flow and Wall Shear Stress in Abdominal Aortic Aneurysms Coupling Laboratory and CFD Simulations
Previous Article in Journal
Laminar Pipe Flow Instability: A Theoretical-Experimental Perspective
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Flow–Structure Interaction and Multiphysics Applications: An Immersed Boundary Perspective

1
Department of Mechanical and Industrial Engineering, Manipal Institute of Technology, Manipal Academy of Higher Education, Manipal 576104, Karnataka, India
2
Department of Aeronautical and Automobile Engineering, Manipal Institute of Technology, Manipal Academy of Higher Education, Manipal 576104, Karnataka, India
3
Department of Biomedical Engineering, Manipal Institute of Technology, Manipal Academy of Higher Education, Manipal 576104, Karnataka, India
4
Department of Mechanical Engineering, St. Joseph Engineering College, Mangaluru 575028, Karnataka, India
*
Author to whom correspondence should be addressed.
Fluids 2025, 10(8), 217; https://doi.org/10.3390/fluids10080217
Submission received: 20 June 2025 / Revised: 25 July 2025 / Accepted: 31 July 2025 / Published: 21 August 2025

Abstract

This article discusses contemporary strategies to deal with immersed boundary (IB) frameworks useful for analyzing flow–structure interaction in complex settings. It focuses on immense advancements in various fields: biology, oscillation of structures due to fluid flow, deformable materials, thermal processes, settling particles, multiphase systems, and sound propagation. The discussion also involves a review of techniques addressing moving boundary conditions at complex interfaces. Evaluating practical examples and theoretical challenges that have been addressed by these frameworks are another focus of the article. Important results highlight the integration of IB methods with adaptive mesh refinement and high-order accuracy techniques, which enormously improve computational efficiency and precision in modeling complex solid–fluid interactions. The article also describes the evolution of IB methodologies in tackling problems of energy harvesting, bio-inspiration propulsion, and thermal-fluid coupling, which extends IB methodologies broadly in many scientific and industrial areas. More importantly, by bringing together different insights and paradigms from across disciplines, the study highlights the emerging trends in IB methodologies towards solving some of the most intricate challenges within the technical and scientific domains.

1. Introduction

This review deals with the historical progression of the immersed boundary method (IBM), tracing its evolution from early theoretical frameworks to the latest developments and diverse applications. With the ability to model complex fluid–structure interactions, IB methods have undergone significant refinements, including high-order accuracy schemes, adaptive mesh refinement, and hybrid approaches elaborately defining computational fluid dynamics with the aid of machine learning algorithms. These schemes helped to introduce different boundary conditions and extend their applicability across various fields ranging from biology to engineering. Presently, simulations of IB models serve to design blood flow and analyze cellular mechanics in biological systems, whereas applications in engineering range from an assessment of vortex-induced vibration (VIV) over flexible body interaction all the way to multiphase fluid dynamic computations.
Over the last two decades, several comprehensive reviews have explored various aspects of the immersed boundary (IB) method, shedding light on its advancements and applications. Iaccarino and Verzicco (2003) [1] examined the role of IB techniques in turbulent flow simulations, emphasizing their effectiveness in handling complex geometries within turbulence modeling. Mittal and Iaccarino (2005) [2] provided a broad overview of IB methods, tracing their evolution and highlighting their significance in fluid dynamics, particularly in scenarios involving moving or deforming boundaries. Sotiropoulos and Yang (2013) [3] focused on the application of IB methods in fluid–structure interaction (FSI), particularly in aerospace engineering, addressing the challenges of accurately modeling interactions between fluids and flexible structures. Kim and Choi (2019) [4] presented a detailed analysis of advancements in IB methods for FSI, encompassing a wide range of applications from biological systems to engineering solutions. Griffith and Patankar (2019) [5] explored recent developments in IB methods for FSI, emphasizing computational efficiency improvements and their application in biological and engineering contexts. Additionally, Roy et al.’s (2020) [6] book offered an extensive discussion on IB methods, covering their theoretical foundations, numerical approaches, and applications across diverse fields, including engineering and biology. Mittal and Bhardwaj (2022) [7] concentrated on the role of IB methods in thermo-fluid simulations, particularly in heat transfer and fluid dynamics. More recently, Verzicco (2022) [8] provided a historical overview of IB methods, tracing their progression over time and discussing future directions, particularly in advancing computational techniques and expanding their applicability in complex fluid dynamics.
Despite these protracted and extensive studies, none of the studies addressed developments post-2020 in a systematic and consolidated manner, especially in terms of comparative analysis of performance across different IB approaches. This review distinctly contributes by analyzing the latest innovations, bridging the gap between practical applications and theoretical insight, and examining methodological advances in a multidisciplinary approach. In contradiction to previous studies, the emphasis here is evolving perspectives on boundary representation, solver stability, and real-time coupling frameworks. In the present review, 140 articles published after 2020 are analyzed. The literature on the immersed boundary method (IBM) highlights a notable gap in the comparative assessment and optimization of sharp versus diffuse boundary treatments in relation to fluid solvers, particularly concerning their performance in biologically relevant flows, flexible body interactions, heat transfer, sound acoustics, and particle sedimentation applications. While both sharp and diffuse boundary treatments have been explored independently in various contexts, there is a lack of systematic studies evaluating their relative strengths and limitations in these specialized domains. For instance, in biological flows and flexible body interactions, the choice between sharp and diffuse interfaces may significantly influence the accuracy of fluid–structure coupling, yet a comprehensive analysis of their effects on deformation dynamics and force transmission remains underexplored. Similarly, in heat transfer problems, the treatment of thermal boundaries—whether sharply resolved or diffusely smoothed—can critically impact temperature gradients and convective heat transfer predictions, but few studies have rigorously compared these approaches. In acoustics, the numerical treatment of boundaries can introduce spurious reflections or damping, yet the optimal boundary representation for sound wave propagation in IBM frameworks has not been thoroughly established. Furthermore, in particle-laden flows and sedimentation studies, the interplay between boundary sharpness and particle–wall interactions—such as adhesion, rebound, or near-wall clustering—lacks detailed investigation, despite its importance in industrial and environmental applications. The absence of generalized guidelines for selecting boundary treatments across these diverse applications limits the broader adoption of IBM, particularly in interdisciplinary research, where accuracy and computational efficiency are paramount. Addressing this gap requires a unified evaluation framework that bridges theoretical insights with practical validation across multiple physical scenarios, ultimately enhancing the reliability and versatility of IBM in complex, real-world simulations.
The present review discusses the immersed boundary method (IBM) in applications like biology, vortex-induced vibration, particle sedimentation, sound acoustics, and multiphase flow. This is important because these areas involve complex fluid–structure interactions and deforming boundaries, where IBM’s ability to handle stability, accuracy, and efficiency enables high-fidelity simulations. Moreover, advancements in IBM have expanded its applicability to cutting-edge fields like bio-inspired propulsion and fluid–acoustic coupling, demonstrating its versatility in addressing real-world challenges across research and industry. These applications highlight IBM’s critical role in modeling intricate systems where traditional methods may struggle, making it indispensable for advancing scientific and engineering solutions. This review therefore emphasizes potential future directions such as multiscale IBM coupling strategies, optimization driven by data of interface models, and robust benchmarking protocols, which are currently lacking in the literature. These contributions aim to improve IBM’s applicability in arriving at solutions for next-generation challenges in engineering and science. By tightly linking the original basis with the latest insights, the review gives a multidisciplinary interpretation, providing researchers and practitioners with the set of tools necessary to employ advanced IB techniques in solving more complex problems in science and engineering.

Literature Search Strategy and Selection Criteria

To ensure and maintain methodological transparency and mitigate selection bias, an organized and systematic literature search was conducted using major academic databases, including Scopus, Elsevier, and Google Scholar. The search encompassed articles published from January 2020 to December 2024, with the primary focus on recent developments in the immersed boundary method (IBM). Keywords used included: immersed boundary method, fluid–structure interaction, diffuse interface, sharp interface, adaptive mesh, bio-inspired propulsion, heat transfer, sound acoustics, particle sedimentation, and combinations thereof.
A total of 482 articles were initially taken into consideration. After elimination of replicas and initial screening based on titles and abstracts, 243 records were evaluated for full-text eligibility. Out of 243, 140 articles were filtered out and were included in the final review based on the following inclusion criteria:
  • Peer-reviewed journal studies or high-impact conference papers.
  • Clear alignment with the advancements in IBM post-2020.
  • Incorporation of original numerical or methodological contributions.
  • Application of IBM in at least one complex fluid–structure or multi-physics context.
Exclusion criteria comprised:
  • Editorials, short communications, or non-technical reviews.
  • Articles focused purely on meshless or ALE methods without relevance to IBM.
  • Studies using IBM terminology vaguely.
The article selection process is illustrated in a PRISMA-style flowchart in Figure 1.

2. Overview of Numerical Framework

A simple overview of the implementation of the IBM is provided. The equations governing the fluid are given by a Navier–Stokes (N–S) equation (Equation (1)) and continuity equation (Equation (2)).
ρ u t + u   ·   u = p + μ 2 u + f
  ·   u = 0
where:
u (x, t) is the fluid velocity (m/s) at position x (m) and time t (s). p (x, t) is pressure in Pa, ρ is density (kg/m3), μ is dynamic viscosity (Pa. s), and f (x, t) is the body force per unit volume, which represents the presence of immersed structure (N/m3).
The immersed structure is often represented as a set of discrete points, typically denoted X (s, t), where s is the Lagrangian coordinate of the structure. The motion of these points is governed by Newton’s laws of motion (Equation (3)),
ρ s 2 X t 2 = F s + F e
where:
The structural density is given by ρ s in (kg/m3), Fs is the restoring elastic force of the structure in (N), and Fe is the external forces, including the fluid force acting on the structure in (N).
The fluid–structure interactions are enforced by velocity interpolation (Equation (4)) and force distribution (Equation (5)).
U s , t = u x , t δ x X s , t d x
f x , t = F s , t δ x X s , t d s
where:
U (s, t) is the Lagrangian velocity (m/s) and F (s, t) is the structural force (N).
For each time step:
Step 1: Compute structural forces F s , t based on elasticity and external loads.
Step 2: Spread structural force to fluid grid: f x , t via Equation (5).
Step 3: Solve Navier–Stokes equations to update u and p using Equations (1) and (2).
Step 4: Interpolate fluid velocity at Lagrangian points using Equation (4).
Step 5: Update structure positions X (s, t) using Newton’s law (Equation (3)).
Step 6: Repeat for the next time step.
Force distribution and velocity interpolation are carried out using the Dirac delta function δ . In the IB method, the structure is modeled using a Lagrangian mesh and the fluid with Eulerian mesh. The N–S equations are solved on a uniform Cartesian grid using finite difference, finite volume, or spectral methods. The IB method is executed in a time-step framework. The fluid velocity and pressure are advanced in time using the N–S equations. The structural forces are determined, and the position of the IB is updated. Coupling is enforced by computing the interaction forces (f) and velocity interpolation (U) between the fluid and structure.
The immersed boundary (IB) method offers significant advantages over techniques such as the arbitrary Lagrangian–Eulerian (ALE) method, particularly for problems involving complex, moving, or deformable boundaries. Unlike ALE, which requires continuous mesh regeneration or deformation to track moving interfaces, the IB method uses a fixed Eulerian grid with Lagrangian markers, eliminating remeshing and simplifying computations. This makes it especially suitable for simulations with large structural deformations or fluid–structure interactions. IB methods are also more adaptable to topological changes, such as merging or splitting of interfaces, which are difficult to handle with ALE. Overall, IB methods provide greater flexibility and computational efficiency (Mittal and Iaccarino, 2005) [2].

Comparison Between Diffuse-Interface and Sharp-Interface IB

A comprehensive comparison between diffuse and sharp-interface methods used in fluid simulations is presented in Table 1 and a boundary interface shown in Figure 2. Diffuse-interface methods represent the interface as a smeared region with finite thickness, using phase-field, level-set, or VOF (volume-of-fluid) techniques (Figure 2a). These methods modify the governing equations (Equation (6)) over a narrow band and apply distributed body forces to handle interfacial dynamics. Kanchan and Maniyeri (2019) [9] demonstrated the accuracy of diffuse-interface immersed boundary method through benchmark studies that compared filament behavior across different orbit regimes—rigid, springy, C-shape, and complex—based on bending rigidity and viscous flow forcing (VFF). The orbit regimes were categorized as: rigid (VFF < 0.16), springy (0.16 < VFF < 1.31), C-shape (1.31 < VFF < 10.67), and complex (VFF > 10.67). The results closely matched the experimental data of Forgacs and Mason (1959) [10] and numerical results of Stockie and Green (1998) [11], validating the method’s ability to accurately capture filament deformation and orbit transitions across a wide range of flow conditions. The diffuse-interface IBM employs a smoothed transition zone between fluid and solid domains, making it particularly well suited for modeling biological phenomena involving large deformations, intricate geometries, and multi-physics coupling. However, they may suffer from slight mass conservation issues.
In contrast, sharp-interface methods treat the interface as a mathematically exact, zero-thickness boundary (Figure 2b), applying sharp jump conditions and discrete forces (Equation (7)). These methods use ghost-cell or cut-cell techniques to explicitly enforce interface conditions, offering better accuracy and mass conservation. While numerically more complex, sharp-interface methods are ideal for problems requiring precise interfacial treatment, such as fluid–structure interaction or multiphase flows with clearly defined boundaries. The sharp-interface immersed boundary (IB) method (Bhardwaj and Mittal, 2012) [12] demonstrates superior performance over ALE-based methods in fluid–structure interaction (FSI), as validated by the Turek–Hron benchmark. It achieves high fidelity with benchmark results, matching the oscillation frequency exactly (f = 0.19) and predicting the plate’s Y-tip displacement within ~11% (0.92 D vs. 0.83 D). The method avoids remeshing by using a fixed Cartesian grid (257 × 129) and accurately captures large, nonlinear deformations using a geometrically nonlinear structural model. It remains stable even at low density ratios (ρs/ρf = 1) and yields a lower mean drag coefficient (3.56 vs. 4.13), indicating reduced unsteady flow forces.
Modified Navier–Stokes equation with body force term:
ρ ϕ u t + u   ·   u = p +   ·   [ μ ϕ u + u T + f i n t e r f a c e
Standard Navier–Stokes equation with sharp-interface enforcement:
ρ u t + u   ·   u =   p + μ 2 u + f
With interface jump condition:
σ · n = f s ,  where [[ · ]] denotes a jump across the interface and f s is surface force.

3. Flow–Structure Interaction Application

3.1. Biophysics and Biological Flows

IBM is extensively applied in biology to model interactions of flexible structures with surrounding fluids, such as blood flow around heart valves or the movement of cells in viscous environments. Thus, it provides the methodology of choice for the simulation of complex physiological processes like cardiovascular function, biofilm formation, and microorganism locomotion by tightly coupling deforming biological structures with incompressible fluid dynamics. Let us now consider the implications of IBM boundary treatments in this application area. Table 2 summarizes recent advancements in IBM boundary treatment with respect to fluid solvers used and applications considered.
Originally developed by Peskin for simulating blood flow in the heart, IBM has since evolved into a versatile tool for studying cardiovascular dynamics, cellular mechanics, microbial locomotion, and tissue-level biomechanics. Its ability to seamlessly handle moving boundaries while maintaining computational efficiency has made it indispensable in biomedical research, enabling high-fidelity simulations of heart valves, bacterial flagella, biofilms, and tumor growth.
In cardiovascular applications, diffuse IBM has enabled unprecedented simulations of heart valve dynamics and blood flow phenomena. Griffith et al. (2019) [5] established an adaptive second-order accurate IBM framework that successfully captured the complex interplay between blood flow and flexible valve leaflets, setting a benchmark for cardiac simulations. This work has been extended by Kaiser et al. (2023) [40], who validated diffuse IBM simulations of heart valve hemodynamics against 4D flow MRI measurements, demonstrating the method’s clinical relevance. For vascular flows, Lampropoulos et al. (2021) [31] and Bourantas et al. (2023) [39] developed efficient finite element implementations of diffuse IBM that accurately simulated blood flow in aneurysms and complex vasculature while maintaining computational tractability for clinically relevant geometries. These implementations highlight diffuse IBM’s ability to handle large deformations and complex boundary motions, making it indispensable for cardiovascular research. Nonetheless, this indispensability should be viewed in context, especially when relative to body-fitting methods such as the finite element method (FEM), which have exhibited strengths in clinical modeling. For instance, Griffith (2011) [49] demonstrated an ALE-based FEM framework for simulating patient-specific vascular mechanics, delivering accurate wall stress distributions vital in aneurysm evaluations. Whilst diffuse IBM is versatile and resilient under moving boundaries and large deformations, it can be prone to interface diffusion, influencing stress localization accuracy (Wang et al., 2012) [50]. In contrast, body-fitted FEM facilitates sharper interface resolution, which is beneficial in contexts like aortic dissection modeling (Kheradvar et al., 2014) [51]. However, IBM possesses a practical edge in dealing with complex moving geometries, like heart valves, with reduced complexity of mesh generation and computational stability (Lee et al., 2020) [52]. Hybrid strategies that integrate immersed fluid solvers with solid mechanics formulated using the finite element method (FEM) have also been put forward to harness the advantages of both paradigms in clinical simulations (Gao et al., 2017) [53].
The study of cellular-scale biological phenomena has particularly benefited from diffuse IBM implementations. Maniyeri et al. (2012) [16] and Maniyeri and Kang (2014) [17] pioneered the application of diffuse IBM to bacterial flagellum dynamics, capturing both propulsion mechanisms and the intricate bundling behavior of multiple flagella. Finite volume method (FVM)-based simulations revealed fundamental insights into microorganism locomotion that would be challenging to obtain experimentally. At smaller scales still, Delong et al. (2014) [26] implemented a fluctuating immersed boundary approach for Brownian dynamic simulations, enabling the study of particle interactions in biologically relevant conditions. Kassen et al. (2022) [34] later extended these capabilities to simulate cell–cell interactions in whole blood, demonstrating diffuse IBM’s versatility across multiple biological length scales.
Flexible filament and vesicle dynamics represent another area where diffuse IBM has made significant contributions. Kim and Lai (2010) [15] and Ong et al. (2021) [28] developed finite difference method (FDM)-based approaches for simulating inextensible vesicles in fluid flow, providing fundamental insights into membrane deformation physics. Kanchan and Maniyeri (2019) [9] conducted a comprehensive series of FVM-based studies on flexible filament dynamics. Various deformations or orbit classes were studied, including buckling and recuperation dynamics, as illustrated in Figure 3, Figure 4 and Figure 5. The physics of the problem involved a “flexible structure which was massless and neutrally buoyant, suspended in a viscous incompressible fluid medium.” The numerical strategy demonstrated here was based on the formal second-order IBM scheme provided in studies conducted by Kanchan and Maniyeri (2019) [9]. Further, this methodology was extended to analyze “asymmetric filament deformation in oscillatory flow” (Kanchan and Maniyeri, 2020) [21], “self-excited oscillation of flexible filament placed in channel flow” (Kanchan and Maniyeri, 2020) [22], and “multiple-filament interaction in shear flow” (Kanchan and Maniyeri, 2020) [23].
These works have collectively advanced understanding of diatom chain dynamics and similar biological systems. Lai and Seol (2022) [38] made advancements to the “immersed boundary method by simulating 3D triangulated vesicles in unsteady Navier–Stokes flow,” as illustrated in Figure 6, Figure 7 and Figure 8. All these illustrate the versatility of IB techniques in accurately modeling FSI across a wide range of applications, encompassing cardiovascular interventions and vesicle dynamics.
The lattice Boltzmann method (LBM) has proven to be particularly synergistic with diffuse IBM for various biological applications. De Rosis (2014) [18] employed LBM-based diffuse IBM to study tandem flapping wings, revealing phase difference effects relevant to insect flight dynamics. Coclite et al. (2020) [27] developed a dynamic IBM framework within LBM that handled both rigid and deformable objects in three-dimensional flows, significantly expanding the method’s applicability to general fluid–structure interaction problems. More recently, Zhu et al. (2022) [37] applied LBM-based diffuse IBM to study tea shoot deformation under negative pressure, demonstrating the method’s potential in agricultural robotic applications beyond traditional biological domains. Meng et al. (2020) [25] made significant contributions to this field through their LBM-based diffuse IBM implementation for studying dendritic growth and motion in convective flows, providing crucial insights into solidification dynamics that bridge material science and fluid mechanics applications.
Specialized numerical treatments have been developed to address challenges in specific biological systems. Fai and Rycroft (2018) [20] implemented a finite element method (FEM)-based diffuse IBM with improved accuracy for thin fluid layers, enabling precise simulation of vesicle migration in narrow channels. Casquero et al. (2021) [30] combined diffuse IBM with B-splines and T-splines for capsule and vesicle dynamic simulations, achieving superior representation of biological membrane mechanics. These specialized implementations demonstrate how diffuse IBM can be adapted to maintain accuracy in challenging scenarios involving thin fluid gaps or complex membrane mechanics.
Hemodynamic applications have driven innovations in diffuse IBM’s handling of pulsatile flows and complex geometries. Mirfendereski and Park (2021) [32] developed a diffuse IBM approach for simulating pulsatile flow in stenotic channels, providing insights into blood flow through constricted arteries. Wang et al. (2020) [54] studied self-propelled flexible plates with Navier slip conditions, contributing to the understanding of bio-inspired propulsion mechanisms. These implementations showcase diffuse IBM’s ability to handle both prescribed and emergent boundary motions in physiological flow conditions.
Recent advances have expanded diffuse IBM’s capabilities through integration with modern computational techniques. Ntetsika and Papadopoulos (2022) [35] combined diffuse IBM simulations of filament dynamics in shear flow with artificial neural network predictions, demonstrating how machine learning can complement traditional computational fluid dynamics. Battista et al. (2018) [19] developed accessible Python and MATLAB implementations of diffuse IBM, lowering the barrier to entry for biological flow simulations. For microscale biological flows, Eldoe et al. (2022) [33] developed an FVM-based diffuse IBM approach to investigate rigid filament interactions in oscillatory flows, advancing the understanding of low-Reynolds-number filament dynamics. Maniyeri (2022) [36] extended the capabilities of diffuse IBM to elastic rod dynamics in fluid flow through an FVM implementation, providing new insights into flagellar motion and similar biological propulsion mechanisms. These developments point toward an increasingly interdisciplinary future for diffuse IBM applications.
In microfluidic and electrochemical applications, diffuse IBM has enabled new insights into complex phenomena. Ladiges et al. (2022) [41] implemented a marker-and-cell (MAC)-based diffuse IBM for simulating electrolytes near physical boundaries, particularly studying induced charge electroosmosis. This work demonstrates how diffuse IBM can be extended beyond traditional fluid–structure interaction problems to include electrochemical effects relevant to lab-on-a-chip devices and similar microsystems.
The finite difference method (FDM) has served as a foundational framework for many biomechanical sharp IBM implementations. Luo et al. (2008) [42] pioneered this approach with their novel IBM for phonation studies, successfully capturing the intricate fluid–structure interactions in vocal fold dynamics. This work demonstrated how sharp IBM could handle the complex geometry and tissue mechanics of human phonation while maintaining computational efficiency. The FDM framework was later extended by Bhardwaj and Mittal (2012) [12] through their coupled IBM–FEM solver, which enabled large-scale simulations of flow-induced deformation in flexible biological structures. This hybrid approach combined the computational efficiency of FDM for fluid dynamics with the accuracy of FEM for structural mechanics, creating a powerful tool for biomechanical simulations.
For traumatic injury modeling, sharp IBM has enabled unprecedented simulations of blast effects on biological tissues. Bhardwaj et al. (2013) [43] applied their FDM–FEM sharp IBM framework to study blast-induced eye deformation, providing critical insights into ocular trauma mechanics. Bailoor et al. (2017) [44] further advanced this capability with their compressible FSI implementation using a ghost-cell method, which accurately captured large deformations during blast–structure interactions. These studies showcased sharp IBM’s ability to handle extreme deformation scenarios while maintaining precise boundary representation—a crucial requirement for reliable injury biomechanics simulations.
Cardiovascular applications have particularly benefited from sharp IBM implementations. Bourantas et al. (2021) [45] developed an FEM-based sharp IBM for simulating blood flow in complex vascular geometries, demonstrating superior accuracy in capturing hemodynamic wall shear stresses. Wang et al. (2022) [47] later implemented a moving sharp IBM within a hybrid FEM framework for heart valve flow simulations, successfully capturing both leaflet dynamics and surrounding flow features. The clinical relevance of these approaches was further demonstrated by Brown et al. (2022) [46], who simulated transcatheter aortic valve replacement (TAVR) devices using patient-specific anatomy in an FDM-based sharp IBM framework. These implementations collectively highlight how sharp IBM has advanced cardiovascular device development and surgical planning.
In cancer research, Singh and Kumar (2023) [48] developed a finite volume method (FVM)-based sharp IBM for tumor morphology modeling, eliminating the need for body-conformal grids while accurately capturing bioheat transfer phenomena. This approach proved particularly valuable for simulating irregular tumor geometries and their interaction with surrounding tissues, demonstrating sharp IBM’s versatility across diverse biomedical applications.

Effect of Fluid Solvers on Boundary Treatment for Biological Applications

The implementation of diffuse IBM across different numerical frameworks reveals important trade-offs between accuracy, computational efficiency, and ease of implementation. Finite difference methods, as employed by Heys et al. (2008) [14] for arthropod sensory systems and Ghosh (2021) [29] for biofilm–fluid interactions, offer straightforward implementation and good performance for many biological flow problems. Finite volume methods provide robust conservation properties advantageous for filament dynamics studies, as demonstrated by Kanchan and Maniyeri’s series of studies. Finite element approaches excel in handling complex geometries and material nonlinearities, making them ideal for vascular simulations like those of Lampropoulos et al. (2021) [31]. Lattice Boltzmann methods, with their natural parallelism and kinetic theory foundation, have proven particularly effective for problems involving complex boundary motions and multiphase effects. The choice of numerical solvers for sharp IBM implementations in biomechanics reflects careful consideration of application-specific requirements. FDM-based approaches, as demonstrated by Luo et al. (2008) [42] and Brown et al. (2022) [46], offer computational efficiency for problems where structured grids are appropriate. FEM implementations like those of Bourantas et al. (2021) [45] and Wang et al. (2022) [47] provide superior geometric flexibility for complex anatomical structures.
Additionally, the computational performance of these solvers is just as critical a consideration. Lattice Boltzmann and finite difference methods, owing to their local stencil operations, demonstrate excellent parallel scalability. For instance, GPU formulations of LBM have exhibited over 85% parallel efficiency in large-scale boundary-resolved simulations (Krüger et al., 2016) [55]. Finite element solvers, whilst facilitating geometric flexibility, are memory-intensive. Nevertheless, unconditionally stable time integration schemes have demonstrated stable convergence with iteration counts growing sublinearly with mesh refinement, as noted in vascular flow simulations (Cueto-Felgueroso et al., 2007) [56]. The finite volume solvers utilized by Kanchan and Maniyeri report showed grid convergence and memory-efficient scaling for filament dynamics. Hybrid schemes such as Bhardwaj and Mittal’s FDM–FEM frameworks capitalize on implicit solvers for structural components whilst ensuring computational efficiency in fluid domains. Whilst not all studies document explicit performance benchmarks, available metrics on memory usage, time to solution, and solver robustness assist in informing the method of choice for a given application and available computational resources.

3.2. Vortex-Induced Vibration and Flexible Body Interaction

“Vortex-induced vibration (VIV)” and the “interaction of flexible structure and fluid flow” are critical issues in the broader field of FSI, extending from engineering to biology. The interactions come about when fluid flow excites the oscillation of flexible or slender structures, such as underwater pipelines and aircraft wings, that exhibit complex dynamic behavior. An understanding of these processes helps optimize designs, predict structural responses, and investigate natural systems, such as fish swimming and plant motion in wind. Table 3 illustrates the recent advancements in IBM implementation with various fluid solvers and boundary treatment for solving VIV and flexible body interaction scenarios.
In bio-inspired propulsion systems, diffuse IBM has enabled unprecedented simulations of aquatic and aerial locomotion mechanisms. The lattice Boltzmann method has proven particularly synergistic with diffuse IBM for problems requiring complex boundary motion and efficient parallelization. Wu et al. (2019) [63] and Yan et al. (2021) [72] employed LBM-based diffuse IBM to study flow-induced vibrations of elliptical and circular cylinders, respectively, demonstrating the method’s effectiveness for bluff body aerodynamics. More recently, Fang et al. (2022) [82] and Huang et al. (2022) [83] extended these capabilities to fish swimming hydrodynamics and fish–turbine interactions, addressing important ecological and engineering challenges in hydropower systems. The natural parallelism of LBM combined with diffuse IBM’s geometric flexibility has enabled these large-scale simulations of complex biological flows.
Kumar et al. (2015) [57] pioneered this application through lattice Boltzmann method (LBM) implementation studying the clap-and-fling mechanism in insect flight, revealing crucial low-Reynolds-number aerodynamics. This work was extended by De Rosis (2015) [58], who examined tandem flapping wings in ground effect, providing insights relevant to micro-aerial vehicle design. For aquatic propulsion, Zhang et al. (2020) [67] developed finite difference method (FDM)-based diffuse IBM frameworks to analyze the hydrodynamics of tuna caudal keels and finlets, respectively, identifying key morphological features contributing to swimming efficiency. These studies collectively demonstrate how diffuse IBM’s ability to handle complex moving boundaries makes it indispensable for bio-inspired design. For fundamental studies of flexible structure interactions, Wang and Tian (2020) [135] developed an innovative FDM–FEM coupled diffuse IBM approach to investigate hydrodynamic interactions between flexible flags in Poiseuille flow.
Energy harvesting applications have similarly benefited from advanced diffuse IBM implementations. Wang et al. (2018) [59] and Li et al. (2018) [60] studied oscillating and pitching hydrofoils for energy extraction using FDM-based diffuse IBM, optimizing motion profiles for maximum power generation. Mazharmanesh et al. (2022) [76] and Mazharmanesh et al. (2023) [87] implemented LBM-based diffuse IBM to investigate inverted piezoelectric flags in tandem and side-by-side configurations, identifying distinct chaotic and symmetric regimes that impact energy harvesting performance. These implementations showcase diffuse IBM’s capability to capture both the fluid dynamics and structural response in coupled energy harvesting systems. Luo et al. (2021) [68] created a reduced-order FSI framework using diffuse IBM that maintained accuracy while significantly decreasing computational cost for biofluid systems.
For vortex-induced vibrations and bluff body flows, finite volume method (FVM) implementations of diffuse IBM have provided valuable insights. Xie et al. (2019) [61] studied vortex-induced vibrations of a cylinder with attached filaments, while Wang and Tian (2019) [136] examined elastic bodies in viscous flows relevant to vegetation and turbine applications. Kasbaoui et al. (2021) [69] implemented a DNS-grade diffuse IBM with moving boundaries to study transitions in swirling von Kármán flow. These FVM-based approaches leverage the method’s strong conservation properties to accurately capture wake dynamics and energy transfer mechanisms in turbulent flows. The RANS approach of Ma et al. (2019) [62] provided practical solutions for industrial-scale turbomachinery problems, and Stival et al. (2022) [86] further developed this capability through their LES–IBM implementation for wind turbine flows, demonstrating how diffuse IBM can be combined with advanced turbulence modeling for renewable energy applications.
Coastal and ocean engineering problems have motivated specialized diffuse IBM implementations. Zhao et al. (2020) [66] developed an FVM-based approach for tsunami-like wave impacts on coastal bridges, addressing critical structural engineering challenges. Luo and Zhang (2023) [88] later implemented diffuse IBM in a σ-coordinate hydrodynamic model for general wave–structure interactions, while Liu et al. (2023) [94] combined LBM-based diffuse IBM with adaptive mesh refinement and volume-of-fluid methods for free-surface flows in ocean engineering. Dong et al. (2021) [70] developed a simplified gas kinetic scheme (SGKS)-based diffuse IBM to investigate chordwise deformation effects on batoid fish locomotion. These implementations demonstrate how diffuse IBM can be adapted to handle the unique challenges of environmental flows with moving boundaries and free surfaces.
Recent advances in computational methods have enabled more sophisticated diffuse IBM implementations for complex FSI problems. Tian et al. (2021) [73] developed a finite element method (FEM)-based diffuse IBM with strict energy conservation properties, providing robust solutions for general FSI applications. Park et al. (2023) [92] implemented a semi-Lagrangian Navier–Stokes solver with reduced diffuse IBM for high-inertia/elasticity FSI, while Monteiro and Mariano (2023) [93] created a Fourier pseudo-spectral diffuse IBM for airfoils and vertical-axis wind turbines. In mixing enhancement applications, Yaswanth and Maniyeri (2022) [75] employed an FVM-based diffuse IBM to study fluid mixing in oscillating lid-driven cavities. These advanced implementations push the boundaries of what is possible in FSI simulation, combining numerical innovation with physical insight.
Drag reduction strategies have been another important application area for diffuse IBM. Zhao et al. (2021) [74] studied micro floating rafts for underwater vehicle drag reduction using FDM-based diffuse IBM, while Mao et al. (2022) [78] and Mao et al. (2023) [96] investigated flexible hairy coatings and buckled filaments for similar purposes. These studies leverage diffuse IBM’s ability to handle complex, deforming surface textures that would be challenging to model with traditional boundary-conforming methods. More recently, Zhang et al. (2024) [95] analyzed a perforated plate–filament system, observing novel momentum flux distribution modes that contribute to drag reduction. Kwon et al. (2020) [108] employed a finite volume method (FVM) to study shallow water flows around cylinders, providing insights into tsunami mitigation strategies. Tsai and Lo (2020) [109] developed a sharp-interface IBM to simulate submerged aquaculture nets, addressing fluid–structure interactions in marine environments. For nonlinear wave dynamics, Tong et al. (2021) [113] proposed a high-fidelity sharp-interface solver to analyze wave–structure interactions relevant to offshore engineering. Hanssen and Greco (2021) [114] introduced an overlapping grid method for efficient simulation of wave–body interactions, enhancing accuracy in free-surface flow modeling. These studies collectively demonstrate the versatility of sharp-interface techniques in tackling complex fluid dynamic problems across engineering disciplines.
Emerging applications continue to push diffuse IBM capabilities in new directions. Dong and Huang (2023) [89] implemented a wall-modeled LES–IBM for batoid fish swimming, revealing previously unseen hairpin vortex structures. Guo and Hou (2023) [90] developed a discrete unified gas kinetic scheme (DUGKS) with slip-condition IBM for autonomous underwater vehicle drag reduction. Wu and Guo (2023) [91] created a free-surface LBM with diffuse IBM for aircraft ditching simulations, incorporating surface tension effects. These cutting-edge applications demonstrate how diffuse IBM continues to evolve to meet new engineering challenges.
Vortex-induced vibration (VIV) studies have been significantly advanced through sharp IBM implementations. Borazjani and Sotiropoulos (2009) [97] pioneered FEM-based sharp IBM simulations of two elastically mounted cylinders at Re = 200, establishing benchmark results for VIV phenomena. This work was extended by Griffith and Leontini (2017) [100], who developed an heuristic model for efficient VIV predictions using an FDM–FEM sharp IBM framework. Recent applications to energy harvesting systems include Badhurshah et al. (2021) [112], Badhurshah et al. (2022) [120], who studied the VIV of cylinders with bistable springs, demonstrating how sharp IBM can capture the complex nonlinear dynamics of these systems. For more complex configurations, Gómez et al. (2022) [119] implemented a sharp IBM in Incompact3D to investigate VIV in four circular cylinders, revealing intricate wake interaction patterns.
The choice of numerical solver for sharp IBM implementations reflects careful consideration of application requirements. Finite difference methods (FDMs) have been widely employed for their simplicity and efficiency in fundamental studies. Khalili et al. (2018) [101] developed a high-order ghost-point FDM sharp IBM for compressible viscous flows, while Xu et al. (2020) [105] and Xu et al. (2023) [122] created high-order FDM implementations for nonlinear water waves and wave loads on marine structures. These approaches demonstrate how FDM-based sharp IBM can achieve excellent accuracy for both fundamental flow physics and engineering applications.
For problems requiring geometric flexibility, finite volume methods (FVMs) have proven effective. Majumdar et al. (2020) [103] used FVM sharp IBM to study dynamic transitions in plunging foil flows, capturing the route to chaos in these systems. Seshadri and De (2021) [110] developed a robust FVM sharp IBM framework for general moving body problems, demonstrating the method’s versatility. The method’s adaptability to industrial flows was further showcased by Zargaran et al. (2023) [130] in rotor–stator mixer simulations (Figure 9). The method was employed for modeling rotor motion, which reduced pre-processing time and enabled the modeling of diverse rotor–stator shapes with different motion patterns, as illustrated in Figure 10 and Figure 11.
Hybrid solver approaches have expanded sharp IBM capabilities for coupled FSI problems. Seo and Mittal (2011) [98] addressed spurious pressure oscillations in moving boundary flows through their FDM–FEM cut-cell sharp IBM. Kundu et al. (2020) [107] developed an FDM–FEM sharp IBM with dynamic under-relaxation that improved stability for general FSI simulations. These hybrid implementations combine the strengths of different numerical methods to address challenging multiscale and multiphysics problems.
In renewable energy applications, sharp IBM has enabled detailed studies of energy harvesting mechanisms. Griffith et al. (2016) [99] examined flow around rotationally oscillating cylinders for energy extraction using FDM–FEM sharp IBM. Khedkar and Bhalla (2022) [117] implemented a model predictive control (MPC) sharp IBM for wave energy converters, demonstrating the method’s potential for renewable energy system optimization. For wind energy applications, Ji et al. (2022) [121] and Yang et al. (2023) [126] developed LES-based sharp IBM frameworks for wind turbine wakes and vertical-axis wind turbine (VAWT) simulations, respectively, providing insights crucial for turbine design and performance prediction.
Free-surface flow simulations have particularly benefited from specialized sharp IBM implementations. Xin et al. (2020) [106] developed a ghost-cell sharp IBM for tank sloshing problems within a constrained interpolation profile (CIR) framework. Robaux and Benoît (2021) [111] created a fully nonlinear potential wave tank using FDM sharp IBM, while Yu et al. (2023) [127] implemented an IB–GHPC method for 2D wave tanks with pressure–velocity separation. These approaches demonstrate how sharp IBM can be adapted to handle the unique challenges of free-surface flows while maintaining interface sharpness.
For industrial and offshore applications, sharp IBM has provided valuable design insights. Song et al. (2022) [118] developed a 3D scour model using RANS sharp IBM for complex coastal structures. Tong et al. (2021) [113] studied nonlinear wave–structure interactions for offshore engineering applications. Industrial applications have equally benefited from the method’s robust handling of complex moving geometries, as evidenced by Joachim et al.’s (2023) [132] parallelized Nitsche’s method for stirred tank simulations (Figure 12), which optimized mixing efficiency while resolving impeller-induced turbulence structures, as shown in Figure 13 and Figure 14, respectively.
Recent advances in sharp IB methodology have focused on improving accuracy and robustness. Kolahdouz et al. (2023) [129] developed an immersed Lagrangian–Eulerian (ILE) sharp IBM with improved volume conservation for large-deformation FSI. The “flexible-body immersed Lagrangian–Eulerian (ILE) scheme” combined immersed boundary (IB) methods with a “Dirichlet–Neumann coupling” strategy to distinguish between fluid and solid subregions with separate momentum equations, as seen in Figure 15 and Figure 16. The proposed approach was validated against computational and experimental FSI benchmarks, demonstrating its robustness and accuracy, as observed in Figure 17.
Li et al. (2023) [131] created a 3D level-set sharp IBM for turbomachinery flows that accurately handled complex blade geometries. Kou and Ferrer (2023) [133] combined volume penalization with selective frequency damping (SFD) to suppress spurious waves in moving boundary simulations. These methodological innovations continue to push the boundaries with sharp IBM.
For flexible structure interactions, several studies have demonstrated sharp IBM’s capabilities. Mishra et al. (2019) [102] studied viscoelastic plates attached to cylinders using FDM–FEM sharp IBM. Sharma et al. (2022) [115] investigated the FIV of cylinders with varying cross sections. Pandey et al. (2023) [123] identified distinct lock-in regimes for flow-induced vibrations of elastic plates, correlating these with optimal energy extraction conditions through high-fidelity simulations that resolved both the fluid vorticity fields and structural mode shapes shown in Figure 18, Figure 19 and Figure 20.
In computational wind engineering, Giannenas et al. (2022) [116] and Narváez et al. (2020) [104] employed Incompact3D with sharp IBM to study wake response and tandem cylinder vibrations, respectively. These spectral element-based implementations demonstrate how sharp IBM can be combined with high-order methods for accurate turbulence simulations.
Emerging applications continue to drive sharp IBM development. Chern et al. (2023) [125] implemented LES sharp IBM to study plasma-actuated dynamic stall control on flapping wings. Recent machine learning integrations by Sundar et al. (2023) [128] coupled physics informed neural networks with IBM. Kanchan et al. (2024) [124] further enhanced predictive capabilities, where artificial neural networks trained on sharp-interface IBM data identified six distinct oscillation regimes for elastic plates in low-Reynolds-number flows, enabling real-time response prediction for adaptive energy harvesting systems (Figure 21, Figure 22 and Figure 23).

Effect of Boundary Treatment on Flexible Body Interaction

The choice of numerical solver for diffuse IBM implementations reflects careful consideration of application-specific requirements. LBM approaches excel for problems with complex boundary motion and parallel scalability needs, as demonstrated by Karimnejad et al. (2022) [77] for particle dynamics and Xiao et al. (2022) [84] for water entry/exit problems. For environmental and geophysical flows, Ai et al. (2021) [71] implemented a diffuse IBM framework capable of predicting internal wave generation and propagation in stratified flows. Their FDM-based approach successfully handled the complex density variations and boundary interactions characteristic of these phenomena, opening new possibilities for oceanographic and atmospheric modeling. FDM implementations offer simplicity and efficiency for fundamental studies of flexible structures, as shown by Chen et al. (2020) [65] for flag flapping and Yu and Yu (2022) [80] for aquatic vegetation. FVMs provide robust conservation properties for turbulent and environmental flows, as exemplified by Jin et al. (2022) [81] for solitary waves and Mi et al. (2022) [85] for submerged nets. Recent advances in multiphase FSI simulations have pushed diffuse IBM capabilities to new levels. Wu and Guo (2023) [91] developed a free-surface LBM with diffuse IBM for aircraft-ditching simulations that incorporated surface tension effects. Recent innovations like Agarwal et al.’s (2024) [134] isogeometric analysis coupling (Figure 24, Figure 25 and Figure 26) and Liu et al.’s (2023) [94] AMR-enhanced free-surface solver point toward future directions with cutting-edge computational techniques to address multiscale, multiphysics challenges across engineering and biological domains.
Sharp interface immersed boundary methods (IBMs) significantly enhance simulations of flexible body interactions by precisely enforcing boundary conditions at fluid–structure interfaces. Unlike diffuse approaches, sharp IBM maintains distinct boundaries, enabling accurate prediction of flow-induced vibrations and large deformations critical for energy harvesting and vortex-induced vibration studies. The method’s exact momentum transfer preserves phase relationships between vortex shedding and structural motion, while its geometric flexibility handles complex morphologies (Sharma et al., 2022) [115]; (Kolahdouz et al., 2023) [129]. For turbulent flows, sharp IBM maintains small-scale features affecting flexible body responses, as demonstrated in wind turbine studies (Ji et al., 2022) [121]; (Yang et al., 2023) [126]. Recent advances combine sharp IBM with data-driven approaches (Sundar et al., 2023) [128]; (Kanchan et al., 2024) [124] and multiphysics applications like plasma flow control (Chern et al., 2023) [125]. The technique’s industrial relevance is shown in mixer simulations, where it reduces numerical diffusion (Zargaran et al., 2023) [130]. By preserving physical fidelity at interfaces, sharp IBM provides unmatched accuracy for analyzing flexible body dynamics across engineering applications.

4. Multiphysics Application

4.1. Heat Transfer

The IB method, or immersed boundary method, is a computational way of simulating fluid flow and heat transfer across complex moving boundaries. By embedding solid structures in a fixed Cartesian grid, the necessity of body-fitted meshes is greatly eradicated, and therefore proper solutions can be provided in applications concerning multiphase flows and thermal interactions. The wide range of its application allows for precise modeling of turbulent flow patterns, energy transport, and heat exchange in particulate systems, and is therefore very important in computational physics and engineering. Table 4 summarizes recent work carried out in the field of heat transfer and its corresponding IBM formulation.
In thermal-fluid systems, diffuse IBM implementations have enabled breakthroughs in conjugate heat transfer simulations by simultaneously handling velocity and temperature boundary conditions. Ren et al. (2013) [138] pioneered this capability through their finite difference method (FDM) implementation incorporating both Dirichlet (velocity) and Neumann (heat flux) corrections, establishing a framework for coupled momentum and energy equations. This approach was extended by Chen et al. (2020) [145] in their lattice Boltzmann method (LBM)-based IB–STLBM, which simplified thermal flow simulations while maintaining accuracy for no-slip and temperature boundary conditions at immersed surfaces. Xu and Choi (2023) [146] further improved the numerical stability of IBM by developing a monolithic immersed boundary projection method (MIBPM) for incompressible flows with thermal transport, ensuring accurate resolution of pressure–velocity–temperature coupling in diffuse IBM frameworks, as shown in Figure 27. The approach involved a two-step approximate lower–upper decomposition technique to decouple momentum and energy equations, incorporating immersed boundary forcing, as noted in Figure 28.
These implementations demonstrate how diffuse IBM’s ability to distribute thermal boundary effects across multiple nodes makes it particularly suitable for problems with complex heat transfer pathways.
Lattice Boltzmann methods have shown remarkable synergy with diffuse IBM for thermal and particulate flow applications, benefiting from their kinetic theory foundation and natural parallelism. Jiang et al. (2022) [142] developed a parallel IB–LBM for fully resolved particle-laden flows that maintained computational efficiency while accurately capturing two-way fluid–particle coupling. Zhang et al. (2016) [151] enhanced this capability through their particulate IBM (PIBM) incorporating the discrete element method (DEM) for thermal particle–fluid interactions, enabling detailed study of heat transfer in granular systems. For microscale applications, Hosseini et al. (2021) [147] implemented a multiple-relaxation-time (MRT) LBM with diffuse IBM to analyze elastic vortex generators in microchannels, demonstrating how the method could enhance mixing in confined geometries. These LBM-based approaches leverage diffuse IBM’s ability to handle complex moving boundaries without remeshing—a crucial advantage for particulate and microfluidic systems.
Energy systems and heat-exchanger applications have driven significant innovations in diffuse IBM. Tong et al. (2020) [140] developed a multiblock LBM–IBM framework for fouling and ash deposition studies in heat exchangers, addressing the challenging combination of particulate deposition and thermal performance degradation. Mazharmanesh et al. (2020) [139] examined energy harvesting from multiple piezoelectric flags using LBM-based diffuse IBM, revealing how tandem and side-by-side configurations affect power extraction efficiency. These implementations showcase diffuse IBM’s versatility in handling both thermal and structural coupling in energy systems, where conventional methods often face difficulties with evolving deposit geometries or oscillating energy harvesters. Tao et al. (2022) [143] simplified IB–LBM for thermal flows while maintaining accurate no-slip and temperature boundary conditions, reducing computational overhead for large-scale simulations.
For rarefied gas flows and microscale heat transfer, specialized diffuse IBM implementations have extended the method’s applicability beyond continuum regimes. The adaptability of diffuse IBM was also evident in studies on rarefied gas flows, where Wang et al. (2023) [150] employed a penalty-based approach to account for velocity and temperature jumps in slip-modeled gas flows. This work provided valuable insights into heat transfer behavior in vacuum environments, space propulsion, and other low-pressure applications (Figure 29, Figure 30 and Figure 31).
Advanced thermal-fluid–structure interaction (TFSI) problems have benefited from improved boundary condition treatments in diffuse IBM implementations. Wu et al. (2023) [148] developed an explicit IBM for TFSI with Neumann boundary conditions (heat flux), enabling precise control of thermal energy transfer at fluid–structure interfaces. Next, Wu et al. (2024) [152] further advanced the field by developing an implicit IBM for TFSI problems with Robin boundary conditions, incorporating combined convection and radiation effects in Figure 32. Their study demonstrated how IBM can bridge the gap between computational efficiency and physical accuracy in multiphysics simulations, as shown in Figure 33.
Chen et al. (2020) [149] implemented a penalty-based diffuse IBM for flexible flags in heated channels, capturing the coupled fluid–structure–thermal dynamics responsible for heat transfer enhancement. These approaches demonstrate how diffuse IBM’s boundary smoothing can be strategically employed to handle challenging multiphysics coupling without sacrificing solution accuracy. The finite element method (FEM) has contributed important diffuse IBM variants for thermal-flow problems. Haeri and Shrimpton (2013) [144] developed an implicit fictitious domain approach for flow and heat transfer past immersed objects, combining FEM’s geometric flexibility with diffuse IBM’s moving boundary capabilities. This implementation proved particularly effective for problems requiring precise thermal boundary condition enforcement on complex geometries, where traditional methods might struggle with mesh generation.
Recent advances have focused on improving computational efficiency and physical fidelity in thermal diffuse IBM implementations. Abaszadeh et al. (2022) [141] extended diffuse IBM to radiative heat transfer in irregular geometries, demonstrating the method’s potential for combined mode heat transfer problems. These developments continue to expand diffuse IBM’s applicability to increasingly complex thermal-fluid phenomena. In particulate flow simulations, diffuse IBM has enabled new insights into suspension dynamics and heat transfer. Jiang et al.’s (2022) [142] fully resolved particle-laden flow simulations and Zhang et al.’s (2016) [151] thermal DEM-coupled approach have provided fundamental understanding of how particles affect heat transfer in fluid systems. These implementations leverage diffuse IBM’s natural handling of moving boundaries to track particle motion and its thermal effects without the computational overhead of conforming mesh updates.
Finite volume methods (FVMs) have proven particularly effective for sharp IBM implementations in thermal systems, combining strong conservation properties with geometric flexibility. Pacheco et al. (2005) [153] established an early FVM-based sharp IBM framework for general heat transfer problems using non-staggered grids with rigorous Dirichlet/Neumann boundary condition enforcement. This approach was extended by Lou et al. (2020) [157] to membrane distillation systems incorporating Robin boundary conditions, demonstrating the method’s versatility for complex thermal boundary conditions in desalination applications. For radiative heat transfer in irregular geometries, Mohammadi and Nassab (2021) [158] developed an FVM–IBM that accurately captured surface-to-surface radiation effects while maintaining sharp-interface treatment. Recent advances by Riahi et al. (2023) [160] and Ahn et al. (2023) [161] have further expanded FVM-based sharp IBM capabilities to compressible flows with heat transfer and phase-change problems involving melting/solidification, respectively, showcasing the method’s adaptability across thermal regimes.
The coupling of sharp IBM with hybrid numerical frameworks has enabled innovative solutions for combined heat transfer problems. Mohammadi and Nassab (2021) [159] developed a novel LBM–FVM–IBM hybrid that leveraged the strengths of both methods for radiative–convective heat transfer simulations, with LBM handling the complex radiation physics and FVM providing robust convection treatment. This hybrid approach maintained sharp-interface resolution while efficiently handling the multiphysics nature of the problem. Similarly, Narváez et al. (2021) [164] implemented a dual IBM approach within the InCompact3D framework for turbulent heat transfer, demonstrating how multiple IBM strategies could be combined to address challenging conjugate heat transfer problems in turbulent flows.
Soti et al. (2015) [154] developed a strongly coupled FDM–FEM sharp IBM solver that accurately captured the interplay between flexible structures and convective heat transfer, enabling optimized design of thermal energy harvesters. Garg et al. (2018) [155] and Garg et al. (2020) [156] extended these capabilities to study thermal buoyancy effects on vortex-induced vibrations (VIVs) and wake-induced vibrations (WIVs), revealing how temperature gradients fundamentally alter bluff body dynamics and energy transfer mechanisms. These implementations demonstrate sharp IBM’s ability to maintain solution fidelity in complex multiphysics environments where thermal, structural, and fluid dynamic effects are tightly coupled.
High-order accurate sharp IBM implementations have significantly advanced thermal flow simulations by reducing grid dependence while maintaining boundary precision. Xia et al. (2014) [168] developed a high-order ghost-cell IBM that reduced required grid points by approximately two-thirds for heat transfer problems while maintaining solution accuracy. Fernandez et al. (2011) [170] implemented a spectral method-based sharp IBM for grid-independent 3D heat conduction analysis, demonstrating exponential convergence for time-dependent boundary problems. These high-order approaches prove particularly valuable for thermal systems where boundary layer resolution is crucial, but computational efficiency remains a concern.
In reacting flows and combustion systems, sharp IBM has enabled precise treatment of thermal boundaries in complex geometries. Ou et al. (2022) [167] developed a directional ghost-cell IBM within their DINO framework that accurately handled thermal boundaries in low-Mach reacting flows, maintaining flame structure fidelity near immersed boundaries. This approach demonstrated how sharp IBM could be adapted to handle the unique challenges of reacting flows where both thermal and species boundaries require precise treatment. Similarly, Cruz and Lamballais (2023) [162] implemented a sharp IBM for high-fidelity conjugate heat transfer in pipe flows with irregular geometries, capturing the subtle interactions between flow physics and thermal transport in complex duct systems.
The treatment of multiphase heat transfer problems has particularly benefited from advanced sharp IBM implementations. Wang et al. (2022) [163] improved direct-forcing IBM for particle-laden flows with heat transfer, enabling accurate resolution of thermal interactions between discrete particles and carrier fluids. Tao et al. (2022) [169] extended these capabilities to discrete unified gas kinetic scheme (DUGKS) simulations, combining the advantages of kinetic theory with sharp-interface treatment for fluid–solid heat transfer problems. For industrial applications involving moving boundaries and transient heat transfer, sharp IBM has provided robust solutions. Shrivastava et al. (2013) [171] developed a level set-based IBM for transient CFD with moving boundaries that maintained sharp-interface treatment throughout complex motion sequences. Zhao and Yan (2022) [166] implemented an enriched IBM for interface-coupled multiphysics problems that improved accuracy for convective heat transfer applications. These approaches demonstrate sharp IBM’s ability to handle the combined challenges of geometric complexity, boundary motion, and thermal transport that characterize many industrial thermal systems. Recent comparative studies have provided valuable insights into sharp IBM performance for thermal applications.
Ménez et al. (2023) [172] conducted a comparative study of volume penalization and IBM for compressible flows with thermal boundary conditions, illustrating the strengths of sharp-interface IBM in handling compressible heat transfer problems. As seen in Figure 34, the study investigated the accuracy and computational cost of both methods and extended the investigation to “three-dimensional supersonic flow configurations using the penalization method.” The instantaneous vorticity structures, iso-surface, and isotherm along with a side view of flapping profile are shown in Figure 35.
Their work demonstrated how sharp IBM implementations could maintain superior accuracy for thermal boundary treatment while providing computational efficiency advantages for certain classes of problems. The lattice Boltzmann method (LBM) has also contributed to sharp IBM advancements in thermal-fluid systems. Wu et al. (2022) [165] developed an explicit IB–RTLBFS method for thermal-fluid–structure interaction problems with Dirichlet boundary conditions, demonstrating how LBM’s kinetic theory foundation could be combined with sharp-interface treatment for efficient thermal simulations. This approach proved particularly effective for problems requiring frequent boundary updates or complex thermal-structural coupling.

4.2. Particle Sedimentation, Multiphase Flows, and Sound Acoustics

The topic of multiphysics applications encompasses areas such as particle sedimentation, multiphase flow, and sound acoustics. The particle sedimentation technique simulates the motion of solid particles dissolved in a fluid under the influence of gravity, dragging, lift, and fluid-induced forces into consideration. The IBM allows for efficient simulation of droplet dynamics, bubble formation, and fluid–structure interactions without requiring deformable meshes. In acoustics, it is used for studying aeroacoustics, wave scattering, and noise control, accurately capturing sound–fluid–structure interactions. This technique finds application in engineering, oceanography, and aerospace to improve the understanding of multiphase transport and the behavior of acoustic waves. Table 5 illustrates recent applications of IBM in multiphysics scenarios.
Zhang et al. (2019) [173] developed a coupled DEM–IB–CLBM framework for erosive particle impacts that accurately captured both particle trajectories and surface wear mechanisms. This approach was extended by Wang et al. (2021) [175] through their polygonal DEM–LBM coupling with an energy-conserving contact algorithm, enabling simulation of arbitrarily shaped particles in viscous flows. For biological suspensions, Yadav and Ghosh (2022) [181] and Panghal and Ghosh (2023) [185] implemented diffuse IBM to study the settling dynamics of permeable planktonic particles, revealing how microscale features influence macroscopic suspension behavior. These implementations demonstrate how diffuse IBM’s ability to handle complex particle geometries and interactions makes it indispensable for suspension rheology studies, as further evidenced by Kawaguchi et al.’s (2022) [182] comparative analysis of IBM and VFM for suspension flows.
Wang et al. (2022) [178] developed an IB–LBM framework for elliptical particle deposition in viscous flows, capturing the complex interplay between particle shape and deposition patterns that influence industrial coating processes. This work complemented Fukui and Kawaguchi’s (2022) [179] investigation of microscopic particle arrangement effects on suspension rheology in narrow channels, where diffuse IBM’s ability to resolve near-wall particle interactions proved crucial for predicting bulk flow properties.
Lattice Boltzmann methods (LBMs) have shown remarkable synergy with diffuse IBMs for particulate and multiphase flow applications, benefiting from their kinetic theory foundation and natural parallelism. Romanus et al. (2021) [177] developed a domain-transferring LBM–IBM for high-Reynolds-number particle settling that maintained accuracy while improving computational efficiency. Cheng and Wachs (2022) [180] implemented an adaptive octree-grid IB–LBM for particle-resolved flows that automatically refined resolution near particles while coarsening elsewhere. Recent advances by Ghosh et al. (2023) [188] produced a stabilized IBM for light particle suspensions (density ratio ≥0.04), addressing a longstanding challenge in buoyant particle simulations. These LBM-based approaches leverage diffuse IBM’s ability to handle complex moving boundaries without remeshing—a crucial advantage for systems with numerous interacting particles.
For biological and flexible particle systems, Ghosh and Panghal (2022) [183] implemented a diffuse IBM to study flexible circular particle settling dynamics, revealing deformation mechanisms that affect sedimentation rates in viscoelastic fluids. This framework was extended by Yadav et al. (2023) [187] to examine semi-torus-shaped permeable particle settling, demonstrating how diffuse IBM can handle both geometric complexity and permeability effects in biological sedimentology applications. These studies collectively advanced understanding of non-spherical particle dynamics in viscous flows, with implications for industrial separations and environmental processes.
In multiphase flow simulations, Patel and Natarajan (2018) [189] developed an interpolation-free IBM that robustly handled moving bodies in multiphase environments, while Niu et al. (2022) [191] presented a simple and effective method for simulating multiphase flows with curved boundaries called the “diffuse-interface immersed-boundary scheme.” A schematic plot of a droplet’s impact on a circular cylinder is shown in Figure 36, and its corresponding shape evolution and velocity field at different time points are shown in Figure 37.
These implementations addressed critical challenges in maintaining interface sharpness while accommodating complex boundary motion. The method’s acoustic applications were advanced by Cheng et al. (2021) [198], who developed a semi-implicit IBM for viscous flow-induced sound generation in computational aeroacoustics. In studies by Hou et al. (2023) [196], a simulation of the entire acoustic field propagation utilizing an immersed-boundary method was carried out, as shown in Figure 38. The study analyzed linear resonance responses for each bubble concerning interior gas velocity. Furthermore, the study delved into a frequency domain analysis of the coupled interactions between the bubbles. Notably, the method accommodated scenarios where two bubbles of different sizes and shapes were in contact with each other, providing insights applicable to underwater scattering targets (Figure 39).
Advanced adaptive implementations have pushed diffuse IBM capabilities further. Zeng et al. (2022) [199] created an adaptive DLM–IBM with subcycling that efficiently handled single and multiphase FSI problems through dynamic resolution control. Yan et al. (2022) [200] (appearing twice in source data) developed an IB–MLBFS with flux correction that improved accuracy for multiphase FSI simulations. These adaptive approaches demonstrate how diffuse IBM continues to evolve to meet the demands of increasingly complex simulations. Hori et al. (2022) [201] contributed a Eulerian-based IBM with implicit lubrication for dense suspensions, addressing the challenging regime of closely interacting particles.
For multiphase flows with moving interfaces, diffuse IBM has enabled new insights into interfacial phenomena and phase-change processes. Zhang et al. (2021) [176] studied coffee-ring formation in evaporating droplets using IB–LBM, capturing both fluid dynamics and deposition patterns. Zhang et al. (2023) [186] extended this to nanofluid droplet freezing with particle expulsion, demonstrating how diffuse IBM can handle solidification fronts and particle interactions simultaneously. In industrial multiphase applications, Souza et al. (2022) [190] combined adaptive mesh refinement (AMR), volume of fluid (VOF), and IBM for turbulent multiphase FSI, showcasing the method’s scalability for complex industrial flows. These implementations highlight diffuse IBM’s versatility in handling multiple interacting physics while maintaining computational tractability.
Bio-inspired applications have driven significant innovations in diffuse IB methodology. Wang and Tian (2019) [136] and Wang and Tian (2020) [135] developed FDM- and FEM-based diffuse IBM frameworks for flapping wing aerodynamics and bioacoustics, revealing how wing flexibility influences both thrust generation and sound production. Wang et al. (2023) [192] created an energy-stable IBM for deformable biological membranes with non-uniform properties, addressing the challenge of simulating heterogeneous elastic structures. These studies demonstrate how diffuse IBM’s ability to handle large deformations and complex material properties makes it invaluable for biomechanical and bio-inspired system studies.
In marine and coastal engineering, diffuse IBM implementations have advanced simulation capabilities for free-surface flows and wave–structure interactions. Ye et al. (2020) [193] developed a discrete-forcing IBM with VOF for ship hydrodynamics that accurately captured wave patterns and hull interactions. Zhang et al. (2013) [205] created a two-phase IB–VOF model specifically for ocean engineering problems, demonstrating robust performance in challenging free-surface conditions. Nangia et al. (2019) [197] implemented a distributed Lagrange multiplier (DLM) IBM with AMR for wave–structure interaction (WSI) problems, enabling efficient simulation of high-density ratio systems. These applications showcase diffuse IBM’s ability to handle the combined challenges of free surfaces, moving structures, and turbulent flows that characterize marine environments.
Acoustic applications have benefited from specialized diffuse IBM formulations. Bilbao (2023) [194] and Bilbao (2023) [195] developed 1D and 3D diffuse IBM frameworks for acoustic wave propagation, enabling efficient simulation of sound barriers and irregular boundaries. Cheng et al. (2017) [204] created a compressible IBM for flow-induced noise using influence matrices, while Wang et al. (2020) [24] implemented a high-order IBM for coupled fluid–structure–acoustics in flapping foils. These approaches demonstrate how diffuse IBM can be adapted to handle the unique requirements of aeroacoustic and computational acoustics problems, where precise wave propagation and boundary interactions are crucial.
The finite difference method (FDM) has served as a versatile foundation for many diffuse IBM implementations. Yang et al. (2019) [174] employed FDM-based DNS with diffuse IBM to study optimal perturbations in particle-laden turbulent channel flows. Liu et al. (2017) [202] combined DNS, IBM, and discrete particle methods (DPMs) for sediment transport studies, capturing particle distribution in turbulent boundary layers. For non-Newtonian suspensions, Fazli et al. (2023) [207] implemented an FDM-based diffuse IBM for yield-pseudoplastic particulate flows, demonstrating the method’s adaptability to complex rheology. The methodology was first validated using Newtonian benchmark cases, after which it was extended to simulate yield-pseudoplastic cases with stationary or moving particles, as illustrated in Figure 40, Figure 41 and Figure 42. The extent of yield is shown in blue and unyielded region is shown in red.
Advanced computational techniques have expanded diffuse IBM capabilities for challenging multiscale problems. Jiang et al. (2022) [142] developed a parallel friction-resolved DNS (FR-DNS) with lubrication model for large-scale suspension simulations. Hori et al. (2022) [201] created a Eulerian-based implicit IBM with lubrication treatment for dense suspensions. Chang et al. (2023) [184] implemented a hybrid peridynamic–Eulerian–IBM for FSI that combined the strengths of multiple numerical approaches. These innovative implementations demonstrate how diffuse IBM continues to evolve to address increasingly complex physical systems across scales. In non-traditional application domains, diffuse IBM has shown surprising versatility. The method’s adaptability was further highlighted by Bürchner et al. (2023) [206], who employed a γ-scaling IBM for full-waveform inversion in non-destructive testing, enabling high-resolution crack detection through computational wavefield analysis. The study introduced a dimensionless scaling function, γ, inspired by fictitious domain methods (Figure 43), to model void regions in the scalar wave equation problem. Three formulations for mono-parameter FWI and one for two-parameter FWI were developed using γ, as shown Figure 44 and Figure 45.
Wang et al. (2017) [203] developed a compressible multiphase FSI framework with an ANCF structural solver for explosive and impact dynamics, showcasing diffuse IBM’s ability to handle extreme deformation scenarios. These unconventional applications highlight the method’s fundamental flexibility and robustness.
In dense particulate flows, sharp IBM implementations have provided unprecedented insights into suspension dynamics and industrial processes. Chéron et al. (2023) [208] developed a hybrid sharp–diffuse IBM (HyBM) for dense particle-laden flows that combined the accuracy of sharp interfaces with the robustness of diffuse methods through non-symmetrical operators. The method utilized a direct-forcing formulation and a regularization of the transfer function between Eulerian grid points and Lagrangian markers to satisfy the no-slip condition at particle surfaces, as shown in Figure 46. The instantaneous velocity field of the random arrays of monodispersed spheres for different numerical resolutions are provided in Figure 47.
For magnetorheological fluids, Ido et al. (2017) [209] implemented a hybrid LBM–IBM–DEM framework that captured the formation of magnetic particle microstructures under external fields, demonstrating sharp IBM’s ability to handle complex multiphysics coupling. These capabilities were further extended to industrial applications by Zeng et al. (2021) [223], who developed a Lagrangian IBM for proppant transport in hydraulic fractures, addressing critical challenges in oil/gas recovery processes.
The lattice Boltzmann method (LBM) has shown remarkable synergy with sharp IBM for particulate and coating applications. Fukui et al. (2018) [210] established a two-way coupling framework for studying particle rotation effects on suspension rheology, revealing how microscopic dynamics influence macroscopic flow properties. Wu and Chen (2022) [212] advanced these capabilities with a 3D IB–LBM for droplet-particle coating processes that accurately resolved the complex interfacial dynamics in spray coating applications. These implementations leverage LBM’s inherent parallelizability while maintaining sharp IBM’s precise boundary treatment, enabling large-scale simulations of industrial particulate systems.
Finite element methods (FEMs) have contributed accurate high-order sharp IBM implementations for challenging particulate flows. Barbeau et al. (2022) [211] developed a high-order FEM–IBM for flow around sphere packings in fixed-bed reactors, demonstrating spectral convergence for these geometrically complex systems. Duprez et al. (2023) [213] implemented a ϕ-FEM approach for Stokes flow around particles that achieved optimal convergence rates in creeping flow regimes. These FEM-based approaches demonstrate how sharp IBM can be combined with advanced spatial discretization to handle particulate systems with complex geometries and tight tolerances for numerical accuracy.
Acoustic applications have driven significant innovations in sharp IB methodology. Zhao et al. (2021) [215] developed a sharp-interface IBM–APE framework for flow-induced noise prediction in cylinder arrays, capturing both the fluid dynamics and acoustic propagation with high fidelity. Xie et al. (2020) [218] created a Cartesian grid method for acoustic scattering using ghost-cell IBM that maintained accuracy while simplifying grid generation. These were complemented by Zhao et al.’s (2022) [219] FDTD–IBM for underwater acoustic scattering, which demonstrated sharp IBM’s versatility across different acoustic modeling paradigms. These implementations showcase how sharp IBM’s precise boundary treatment is crucial for accurate wave propagation and scattering simulations.
Industrial applications have motivated the development of robust sharp IBM implementations. Gorges et al. (2024) [216] conducted a comprehensive comparison of smooth and blocked-off IBM for fixed-bed reactors, providing guidelines for method selection in industrial packed bed simulations. The comparison was conducted at particle Reynolds numbers of 300 and 500, with “inline particle image velocimetry (PIV)” measurements serving as a basis for evaluation (Figure 48). Both IBM approaches exhibited excellent agreement with the experimental results, especially at higher resolutions, capturing the more stable flow at a particle Reynolds number of 300. However, differences between the IBM approaches arose in the more unsteady flow at a particle Reynolds number of 500. Contour plots of the y-component of the mean velocity vector field and qualitative cutout of the simulation mesh for the smooth IBM are illustrated in Figure 49 and Figure 50, respectively.
Isoz et al. (2022) [220] developed a hybrid fictitious domain–IBM for irregular particle flows commonly encountered in industrial processes. These implementations address the dual challenges of geometric complexity and computational efficiency that characterize many industrial applications.
The finite volume method (FVM) has proven particularly effective for industrial-scale sharp IBM implementations. Giahi and Bergstrom (2023) [222] aimed to verify and validate the immersed boundary methods (IBMs) implemented in FOAM-Extend open-source toolbox versions 4.0 and 4.1. The study comprised five test cases of increasing complexity, wherein simulation results were compared with experimental and numerical data from the literature (Figure 51, Figure 52 and Figure 53).
Zhou and Balachandar (2021) [217] analyzed spatiotemporal resolution requirements in IBM with direct forcing, providing fundamental insights for particulate flow simulations. These studies establish best practices for applying sharp IBM to industrial fluid–particle systems.
Advanced hybrid approaches have expanded sharp IBM’s capabilities for complex boundary interactions. Qin et al. (2022) [221] developed a hybrid IBM for particle–complex boundary interactions that maintained accuracy while handling intricate geometric features. Farooq et al. (2025) [214] implemented a radial-basis function approach within a QUICK scheme for bio-inspired moving bodies, demonstrating sharp IBM’s adaptability to different numerical frameworks. These hybrid implementations combine the strengths of multiple numerical techniques to address challenging fluid–structure interaction problems.
In coastal and geophysical applications, sharp IBM has enabled new simulation capabilities. Moreover, the method’s success in handling free-surface flows, as demonstrated by Yu et al. (2023) [127] in their two-phase wave tank simulations, underscores its flexibility in coastal and offshore engineering applications (Figure 54, Figure 55 and Figure 56).
Caunt et al. (2023) [224] implemented a Taylor-series based IBM for seismic/acoustic wave propagation over irregular terrain, demonstrating the method’s potential for geophysical simulations. These applications showcase sharp IBM’s versatility across different physical scales and regimes.
Theoretical foundations and comparative studies have played a crucial role in advancing sharp IB methodology. Haeri and Shrimpton (2012) [227] provided a comprehensive review of IBM and fictitious domain methods for particulate flows, establishing benchmarks for method selection and implementation. This theoretical framework informs contemporary developments in sharp IBM, ensuring rigorous treatment of fundamental numerical aspects.

4.3. Selection of Fluid Solver and Boundary Condition for Multiphysics Applications

The choice of numerical solver for diffuse IBM implementations in thermal-fluid systems reflects careful consideration of physical requirements and computational constraints. LBM approaches, as demonstrated by Mazharmanesh et al. (2020) [139] and Tong et al. (2020) [140], excel in parallel scalability and complex boundary handling for particulate and energy systems. LBM approaches also excel for particulate and multiphase systems requiring efficient handling of complex boundaries (Zhang et al., 2019; Cheng and Wachs, 2022) [173,180]. FDM implementation offers simplicity and efficiency for fundamental studies of heat transfer enhancement. FEM methods provide geometric flexibility for complex conjugate heat transfer problems, exemplified by Haeri and Shrimpton (2013) [144] and Wang and Tian (2020) [135]. Each framework brings unique advantages that diffuse IBM enhances through its boundary treatment flexibility. Emerging applications continue to push diffuse IBM capabilities in new directions. The method’s extension to radiative heat transfer by Abaszadeh et al. (2022) [141] and to rarefied gas flows by Wang et al. (2023) [150] demonstrates its adaptability to diverse physical regimes. Coupling with DEM for particle-laden flows, as shown by Zhang et al. (2016) [151], illustrates how diffuse IBM can integrate with other numerical techniques to handle multiphysics challenges. These cutting-edge applications suggest a bright future for diffuse IBM in increasingly complex thermal-fluid systems.
The selection of numerical solvers for sharp-interface immersed boundary methods (IBM) in multiphysics applications requires careful consideration of accuracy, efficiency, and physical fidelity, with different solvers offering distinct advantages for specific applications. Lattice Boltzmann methods (LBMs) excel in particulate flows due to their natural handling of complex boundaries and parallelizability (Fukui et al., 2018; Wu and Chen, 2022) [210,212], while finite difference methods (FDMs) and finite-difference time-domain (FDTD) approaches are preferred for acoustic applications requiring precise wave propagation (Zhao et al., 2021; Zhao et al., 2022) [215,219]. Finite element methods (FEMs) provide geometric flexibility for industrial packed-bed reactors (Barbeau et al., 2022) [211], and finite volume methods (FVMs) remain crucial for robust conservation in industrial applications (Giahi and Bergstrom, 2023) [222]. Emerging hybrid approaches like Chéron et al.’s (2023) [208] HyBM and Farooq et al.’s (2025) [214] radial-basis–QUICK combination strategically blend solver strengths to address limitations of individual methods, guided by resolution analyses like those of Zhou and Balachandar (2021) [217]. Current trends show increasing specialization based on dominant physics, scale requirements, and interface complexity, with comparative studies (Haeri and Shrimpton, 2012) [227] informing a pragmatic “horses for courses” approach that matches solvers to specific application needs while employing hybrid methods to bridge gaps, ensuring sharp IBM remains effective for increasingly complex multi-physics challenges.

5. Evolution of IBM

The immersed boundary (IB) method has seen extensive application across multiple domains, including biological systems, vortex-induced vibration (VIV), fluid–structure interactions (FSIs), and engineering simulations. Since 2007, notable advancements have emerged, beginning with the formulation of an adaptive IB scheme for cardiac mechanics, which exhibited second-order accuracy in simulating blood flow around heart valves. In the ensuing years, research concentrated on vesicle–fluid interactions, bacterial flagellar dynamics, and biofilm behavior, facilitating applications in modeling stress and deformation in biological entities, such as eye blast overpressure effects and plankton settling in aquatic ecosystems. Between 2010 and 2012, refinements to IB methodologies were made to investigate hair–fluid interactions employing penalty IB techniques (modified version of immersed boundary method that considers mass of boundary), alongside extensions to model flexible structures in maritime applications. From 2014 to 2017, innovative IB-based FSI solvers were developed for compressible fluid dynamics, thin fluid layers, and vortex interactions, thereby enhancing the understanding of the impact of hydrodynamic forces on vibration amplitudes and energy harvesting mechanisms in structures such as hydrofoils and flags. From 2018 to 2020, IB methods were adeptly adapted for wave–structure interactions, underwater vehicle dynamics, and coastal infrastructure applications, with investigations illustrating their capability to simulate flow-induced vibrations in wind turbines and facilitate energy extraction from wave-induced motions through model predictive control (MPC) techniques. By 2021–2023, research had pivoted towards hybrid IB models that integrated lattice Boltzmann methods and artificial intelligence-driven methodologies to augment numerical precision, turbulence representation, and complex solid–fluid interactions. Recent developments in 2024 focused on IB applications pertaining to highly flexible structures, real-time engineering simulations, and fluid–structure interaction challenges involving intricate geometries. Throughout these advancements, significant innovations have encompassed adaptive mesh refinement, high-order accuracy methodologies, and novel IB strategies aimed at addressing real-world challenges across biological, mechanical, and marine engineering domains.

6. Future Scope and Concluding Remarks

Immersed boundary methods (IBMs) have emerged as indispensable computational tools across diverse fields, with their selection—sharp or diffuse interface—dictated by application-specific demands. In biological fluid–structure interaction (FSI), sharp-interface IBM excels in capturing precise boundary conditions for vocal fold dynamics, cardiovascular flows, and cellular mechanics, while diffuse-interface methods prove advantageous for complex geometries, large deformations, and multi-scale phenomena, despite ongoing challenges in turbulent and high-Reynolds-number flows. For vortex-induced vibrations (VIVs) and flexible body interactions, sharp-interface IBM enables high-fidelity analysis of wake dynamics and industrial applications, whereas diffuse methods offer robustness in handling bio-inspired propulsion and renewable energy systems. In heat transfer, sharp-interface techniques provide accuracy for conjugate, radiative, and multiphase thermal problems, while diffuse methods facilitate simulations of moving boundaries and microscale systems. The integration of IBM with emerging technologies—such as machine learning for adaptive boundary treatments, high-performance computing for large-scale systems, and multiphysics coupling for energy and biomedical applications—highlights its evolving potential. As computational capabilities advance, both sharp and diffuse IBM are poised to address increasingly complex challenges, from turbulent flows and non-Newtonian suspensions to aeroacoustics and microelectronics cooling, solidifying their role as transformative tools in computational science and engineering.
The latest studies indicate several solid strategies that can enhance immersed boundary (IB) frameworks to outperform their current caliber. One potentially effective direction is the incorporation of IB methods with physics-informed neural networks (PINNs), as established by Farea et al. (2025) [228], who proposed an adaptive B-spline PINN architecture for fluid–structure interaction challenges. Their approach, which mitigates mesh deformation whilst resolving complicated boundary behavior, allows accurate modeling of dynamic interfaces such as bioinspired motion and vortex shedding. For high-Reynolds-number flows, Zuber et al. (2025) [229] successfully integrated wall-modeled large-eddy simulation (WMLES) with an immersed boundary solver to simulate full-scale rotorcraft aerodynamics, obtaining realistic vortex dynamics utilizing GPU acceleration. This evaluates the use of hybrid IB–turbulence models in aerospace and energy applications where wall-resolved LES is impractical. Simultaneously, Gopakumar et al. (2025) [230] introduced a conformal prediction framework for uncertainty quantification in neural PDE solvers, which can be expanded to IB–PINN formulations to facilitate statistically sound error bounds without necessitating labeled data—specifically valuable in environmental and biomedical modeling. Lastly, learning-based interfacial flux modeling is rising as a robust and versatile substitute to physics-based closures. The features-embedded learning immersed boundary (FEL-IB) model (Zhou et al., 2025) [231] makes use of neural networks to estimate interfacial momentum exchange, providing enhanced fidelity in multiphase turbulence and thermofluid challenges. Collectively, these techniques mark a shift toward intelligent, scalable, and reliable IB simulations grounded in data-driven and hybrid computational physics. A more systematic discussion of the above can also be found in our previous review (Powar et al., 2025) [232].
The potential for advancing immersed boundary (IB) frameworks lies in addressing key challenges and exploring emerging opportunities. Future research could focus on creating more efficient numerical techniques to handle large-scale simulations involving intricate geometries and dynamically evolving interfaces. Enhancing the precision of these methods for accurately capturing complex behaviors, such as turbulent flows, deformable structures, and multiphase systems, is another critical area of exploration. The integration of IB approaches with machine learning and data-driven strategies could lead to real-time simulation capabilities and predictive analytics. Furthermore, leveraging advances in high-performance computing, such as quantum and exascale platforms, can significantly improve the scalability and resolution of simulations, expanding the scope of applications in fields like energy, medicine, and environmental science.
Alongside methodological and computational advancements, the experimental validation and reproducibility of IB methods remain inadequately addressed, yet are indispensable for the community. Amidst the rising interest in IBM across multiple disciplines, a significant share of the literature is devoid of rigorous grid-independence studies, open-source code repositories, or validation against experimental benchmarks. These exclusions curb the ability to verify results, establish standardized comparisons across different modeling frameworks, and reproduce studies. Future work should primarily focus on the inclusion of these reproducibility metrics and ensure transparency in model development and simulation protocols. Doing so would influentially improve the robustness, credibility, and translational potential of IB research.
Building on the foundational discussions of earlier work, this article has examined the latest progress and implementations of IB methods across disciplines, including biology, VIV, heat transfer, particle interactions, multi-fluid systems, and acoustic phenomena. By addressing theoretical advancements and practical challenges, it highlights the evolving role of IB methodologies in solving sophisticated interaction problems. As the field continues to innovate and expand, IB techniques are poised to play a pivotal role in advancing computational tools and addressing critical challenges across a broad spectrum of scientific and engineering domains.

Funding

This research received no external funding.

Data Availability Statement

All necessary data will be available upon request.

Acknowledgments

The authors deeply acknowledge the support of Rudra Kumar Pandey, Varun Vinod Nair, and Sriniket Srinivasan in providing resources for this project. The authors would also like to acknowledge the support given by way of computing facilities by the Department of Mechanical and Industrial Engineering, Manipal Institute of Technology, Manipal Academy of Higher Education, Manipal.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Iaccarino, G.; Verzicco, R. Immersed boundary technique for turbulent flow simulations. Appl. Mech. Rev. 2003, 56, 331–347. [Google Scholar] [CrossRef]
  2. Mittal, R.; Iaccarino, G. Immersed Boundary Methods. Annu. Rev. Fluid Mech. 2005, 37, 239–261. [Google Scholar] [CrossRef]
  3. Sotiropoulos, F.; Yang, X. Immersed boundary methods for simulating fluid–structure interaction. Prog. Aerosp. Sci. 2013, 65, 1–21. [Google Scholar] [CrossRef]
  4. Kim, W.; Choi, H. Immersed boundary methods for fluid-structure interaction: A review. Int. J. Heat Fluid Flow 2019, 75, 301–309. [Google Scholar] [CrossRef]
  5. Griffith, B.E.; Patankar, N.A. Immersed Methods for Fluid–Structure Interaction. Annu. Rev. Fluid Mech. 2019, 52, 421–448. [Google Scholar] [CrossRef] [PubMed]
  6. Roy, S.; De, A.; Balaras, E. Immersed Boundary Method; Springer: Singapore, 2020. [Google Scholar] [CrossRef]
  7. Mittal, R.; Bhardwaj, R. Immersed Boundary Methods For Thermofluids Problems. Annu. Rev. Heat Transf. 2022, 24, 33–70. [Google Scholar] [CrossRef]
  8. Verzicco, R. Immersed Boundary Methods: Historical perspective and Future Outlook. Annu. Rev. Fluid Mech. 2022, 55, 129–155. [Google Scholar] [CrossRef]
  9. Kanchan, M.; Maniyeri, R. Numerical analysis of the buckling and recuperation dynamics of flexible filament using an immersed boundary framework. Int. J. Heat Fluid Flow 2019, 77, 256–277. [Google Scholar] [CrossRef]
  10. Forgacs, O.L.; Mason, S.G. Particle motions in sheared suspensions: X. Orbits of flexible threadlike particles. J. Colloid Sci. 1959, 14, 473–491. [Google Scholar] [CrossRef]
  11. Stockie, J.M.; Green, S.I. Simulating the motion of flexible pulp fibres using the immersed boundary method. J. Comput. Phys. 1998, 147, 147–165. [Google Scholar] [CrossRef]
  12. Bhardwaj, R.; Mittal, R. Benchmarking a coupled Immersed-Boundary-Finite-Element Solver for Large-Scale Flow-Induced deformation. AIAA J. 2012, 50, 1638–1642. [Google Scholar] [CrossRef]
  13. Griffith, B.E.; Hornung, R.D.; McQueen, D.M.; Peskin, C.S. An adaptive, formally second-order accurate version of the immersed boundary method. J. Comput. Phys. 2007, 223, 10–49. [Google Scholar] [CrossRef]
  14. Heys, J.J.; Gedeon, T.; Knott, B.C.; Kim, Y. Modelling arthropod filiform hair motion using the penalty immersed boundary method. J. Biomech. 2008, 41, 977–984. [Google Scholar] [CrossRef]
  15. Kim, Y.; Lai, M. Simulating the dynamics of inextensible vesicles by the penalty immersed boundary method. J. Comput. Phys. 2010, 229, 4840–4853. [Google Scholar] [CrossRef]
  16. Maniyeri, R.; Suh, Y.K.; Kang, S.; Kim, M.J. Numerical study on the propulsion of a bacterial flagellum in a viscous fluid using an immersed boundary method. Comput. Fluids 2012, 62, 13–24. [Google Scholar] [CrossRef]
  17. Maniyeri, R.; Kang, S. Numerical study on bacterial flagellar bundling and tumbling in a viscous fluid using an immersed boundary method. Appl. Math. Model. 2014, 38, 3567–3590. [Google Scholar] [CrossRef]
  18. De Rosis, A. On the dynamics of a tandem of asynchronous flapping wings: Lattice Boltzmann-immersed boundary simulations. Phys. A Stat. Mech. Appl. 2014, 410, 276–286. [Google Scholar] [CrossRef]
  19. Battista, N.A.; Strickland, W.C.; Barrett, A.; Miller, L.A. IB2d Reloaded: A more powerful Python and MATLAB implementation of the immersed boundary method. Math. Methods Appl. Sci. 2018, 41, 8455–8480. [Google Scholar] [CrossRef]
  20. Fai, T.G.; Rycroft, C.H. Lubricated immersed boundary method in two dimensions. J. Comput. Phys. 2018, 356, 319–339. [Google Scholar] [CrossRef]
  21. Kanchan, M.; Maniyeri, R. Numerical simulation of buckling and asymmetric behaviour of flexible filament using temporal second-order immersed boundary method. Int. J. Numer. Methods Heat Fluid Flow 2020, 30, 1047–1095. [Google Scholar] [CrossRef]
  22. Kanchan, M.; Maniyeri, R. Fluid-structure interaction study and flowrate prediction past a flexible membrane using immersed boundary method and artificial neural network techniques. J. Fluids Eng. 2020, 142, 051501. [Google Scholar] [CrossRef]
  23. Kanchan, M.; Maniyeri, R. Numerical simulation and prediction model development of multiple flexible filaments in viscous shear flow using immersed boundary method and artificial neural network techniques. Fluid Dyn. Res. 2020, 52, 045507. [Google Scholar] [CrossRef]
  24. Wang, L.; Tian, F.; Lai, J. An immersed boundary method for fluid–structure–acoustics interactions involving large deformations and complex geometries. J. Fluids Struct. 2020, 95, 102993. [Google Scholar] [CrossRef]
  25. Meng, S.; Zhang, A.; Guo, Z.; Wang, Q. Phase-field-lattice Boltzmann simulation of dendrite motion using an immersed boundary method. Comput. Mater. Sci. 2020, 184, 109784. [Google Scholar] [CrossRef]
  26. Delong, S.; Usabiaga, F.B.; Delgado-Buscalioni, R.; Griffith, B.E.; Donev, A. Brownian dynamics without Green’s functions. J. Chem. Phys. 2014, 140, 134110. [Google Scholar] [CrossRef]
  27. Coclite, A.; Ranaldo, S.; Pascazio, G.; De Tullio, M.D. A Lattice Boltzmann dynamic-Immersed Boundary scheme for the transport of deformable inertial capsules in low-Re flows. Comput. Math. Appl. 2020, 80, 2860–2876. [Google Scholar] [CrossRef]
  28. Ong, K.C.; Lai, M.; Seol, Y. An immersed boundary projection method for incompressible interface simulations in 3D flows. J. Comput. Phys. 2021, 430, 110090. [Google Scholar] [CrossRef]
  29. Ghosh, S. Immersed boundary simulations of fluid shear-induced deformation of a cantilever beam. Math. Comput. Simul. 2021, 185, 384–402. [Google Scholar] [CrossRef]
  30. Casquero, H.; Bona-Casas, C.; Toshniwal, D.; Hughes, T.J.; Gómez, H.; Zhang, Y. The divergence-conforming immersed boundary method: Application to vesicle and capsule dynamics. J. Comput. Phys. 2021, 425, 109872. [Google Scholar] [CrossRef]
  31. Lampropoulos, D.S.; Bourantas, G.C.; Zwick, B.F.; Kagadis, G.C.; Wittek, A.; Miller, K.; Loukopoulos, V.C. Simulation of intracranial hemodynamics by an efficient and accurate immersed boundary scheme. Int. J. Numer. Methods Biomed. Eng. 2021, 37, e3524. [Google Scholar] [CrossRef]
  32. Mirfendereski, S.; Park, J.S. Direct numerical simulation of a pulsatile flow in a stenotic channel using immersed boundary method. Eng. Rep. 2021, 4, e12444. [Google Scholar] [CrossRef]
  33. Eldoe, J.B.; Kanchan, M.; Maniyeri, R. Modelling rigid filament interaction under oscillatory flow using immersed boundary method. Mater. Today Proc. 2022, 56, 785–790. [Google Scholar] [CrossRef]
  34. Kassen, A.; Barrett, A.; Shankar, V.; Fogelson, A.L. Immersed boundary simulations of cell-cell interactions in whole blood. J. Comput. Phys. 2022, 469, 111499. [Google Scholar] [CrossRef]
  35. Ntetsika, M.; Papadopoulos, P. Numerical simulation and predictive modelling of an inextensible filament in two-dimensional viscous shear flow using the Immersed Boundary/Coarse-Graining Method and Artificial Neural Networks. Comput. Methods Appl. Mech. Eng. 2022, 401, 115589. [Google Scholar] [CrossRef]
  36. Maniyeri, R. Numerical modelling of straight and helical elastic rods under fluid flow using immersed boundary method. Mater. Today Proc. 2022, 56, 686–689. [Google Scholar] [CrossRef]
  37. Zhu, Y.; Wei, Y.; Wang, Z.; Wang, R.; Wu, C.; Chen, J.; Tong, J. Numerical simulation for deformation characteristic of tea shoot under negative pressure guidance by the immersed boundary–lattice Boltzmann method. J. Comput. Sci. 2022, 65, 101882. [Google Scholar] [CrossRef]
  38. Lai, M.; Seol, Y. A stable and accurate immersed boundary method for simulating vesicle dynamics via spherical harmonics. J. Comput. Phys. 2022, 449, 110785. [Google Scholar] [CrossRef]
  39. Bourantas, G.C.; Zwick, B.F.; Lampropoulos, D.S.; Loukopoulos, V.C.; Κατσάνος, Κ.; Dimas, A.A.; Burganos, V.N.; Wittek, A.; Miller, K. A voxelized immersed boundary (VIB) finite element method for accurate and efficient blood flow simulation. arXiv 2023, arXiv:2007.02082. [Google Scholar] [CrossRef]
  40. Kaiser, A.D.; Schiavone, N.; Elkins, C.J.; McElhinney, D.B.; Eaton, J.K.; Marsden, A.L. Comparison of immersed boundary simulations of heart valve hemodynamics against in vitro 4D flow MRI data. Ann. Biomed. Eng. 2023, 51, 2267–2288. [Google Scholar] [CrossRef]
  41. Ladiges, D.R.; Wang, J.G.; Srivastava, I.; Nonaka, A.; Bell, J.B.; Carney, S.P.; Garcia, A.L.; Donev, A. Modeling electrokinetic flows with the discrete ion stochastic continuum overdamped solvent algorithm. Phys. Rev. E 2022, 106, 035104. [Google Scholar] [CrossRef] [PubMed]
  42. Luo, H.L.; Mittal, R.; Zheng, X.; Bielamowicz, S.; Walsh, R.J.; Hahn, J.K. An immersed-boundary method for flow–structure interaction in biological systems with application to phonation. J. Comput. Phys. 2008, 227, 9303–9332. [Google Scholar] [CrossRef]
  43. Bhardwaj, R.; Ziegler, K.; Seo, J.H.; Ramesh, K.; Nguyen, T.D. A computational model of blast loading on the human eye. Biomech. Model. Mechanobiol. 2013, 13, 123–140. [Google Scholar] [CrossRef]
  44. Bailoor, S.; Annangi, A.; Seo, J.H.; Bhardwaj, R. Fluid-structure interaction solver for compressible flows with applications to blast loading on thin elastic structures. Appl. Math. Model. 2017, 52, 470–492. [Google Scholar] [CrossRef]
  45. Bourantas, G.C.; Lampropoulos, D.S.; Zwick, B.F.; Loukopoulos, V.C.; Wittek, A.; Miller, K. Immersed boundary finite element method for blood flow simulation. Comput. Fluids 2021, 230, 105162. [Google Scholar] [CrossRef]
  46. Brown, J.; Lee, J.H.; Smith, M.; Wells, D.; Barrett, A.; Puelz, C.; Vavalle, J.P.; Griffith, B.E. Patient–Specific immersed Finite Element–Difference model of transcatheter aorticvalve replacement. Ann. Biomed. Eng. 2022, 51, 103–116. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, J.; Zhou, C.; Ai, J. A hybrid immersed-boundary/body-fitted-grid method and its application to simulating heart valve flows. Int. J. Numer. Methods Fluids 2022, 94, 1996–2019. [Google Scholar] [CrossRef]
  48. Singh, A.; Kumar, N. A coupled finite-volume immersed boundary method for the simulation of bioheat transfer in the 3D complex tumour. Eng. Comput. 2023, 39, 3743–3758. [Google Scholar] [CrossRef]
  49. Griffith, B.E. Immersed boundary model of aortic heart valve dynamics with physiological driving and loading conditions. Int. J. Numer. Methods Biomed. Eng. 2011, 28, 317–345. [Google Scholar] [CrossRef]
  50. Wang, H.M.; Gao, H.; Luo, X.Y.; Berry, C.; Griffith, B.E.; Ogden, R.W.; Wang, T.J. Structure-based finite strain modelling of the human left ventricle in diastole. Int. J. Numer. Methods Biomed. Eng. 2012, 29, 83–103. [Google Scholar] [CrossRef]
  51. Kheradvar, A.; Groves, E.M.; Dasi, L.P.; Alavi, S.H.; Tranquillo, R.; Grande-Allen, K.J.; Simmons, C.A.; Griffith, B.; Falahatpisheh, A.; Goergen, C.J.; et al. Emerging Trends in heart valve Engineering: Part I. Solutions for future. Ann. Biomed. Eng. 2014, 43, 833–843. [Google Scholar] [CrossRef]
  52. Lee, J.H.; Rygg, A.D.; Kolahdouz, E.M.; Rossi, S.; Retta, S.M.; Duraiswamy, N.; Scotten, L.N.; Craven, B.A.; Griffith, B.E. Fluid–Structure interaction models of bioprosthetic heart valve dynamics in an experimental pulse duplicator. Ann. Biomed. Eng. 2020, 48, 1475–1490. [Google Scholar] [CrossRef]
  53. Gao, H.; Feng, L.; Qi, N.; Berry, C.; Griffith, B.E.; Luo, X. A coupled mitral valve—Left ventricle model with fluid–structure interaction. Med. Eng. Phys. 2017, 47, 128–136. [Google Scholar] [CrossRef]
  54. Wang, S.; Ryu, J.; He, G.; Qin, F.; Sung, H.J. A self-propelled flexible plate with a Navier slip surface. Phys. Fluids 2020, 32, 021906. [Google Scholar] [CrossRef]
  55. Krüger, T.; Kusumaatmaja, H.; Kuzmin, A.; Shardt, O.; Silva, G.; Viggen, E.M. The Lattice Boltzmann Method; Graduate Texts in Physics; Springer: Berlin/Heidelberg, Germany, 2016. [Google Scholar] [CrossRef]
  56. Cueto-Felgueroso, L.; Colominas, I.; Nogueira, X.; Navarrina, F.; Casteleiro, M. Finite volume solvers and Moving Least-Squares approximations for the compressible Navier–Stokes equations on unstructured grids. Comput. Methods Appl. Mech. Eng. 2007, 196, 4712–4736. [Google Scholar] [CrossRef]
  57. Kumar, S.; De, A.; Das, D. Investigation of the flow field of clap and fling motion using immersed boundary coupled lattice Boltzmann method. J. Fluids Struct. 2015, 57, 247–263. [Google Scholar] [CrossRef]
  58. De Rosis, A. Ground-induced lift enhancement in a tandem of symmetric flapping wings: Lattice Boltzmann-immersed boundary simulations. Comput. Struct. 2015, 153, 230–238. [Google Scholar] [CrossRef]
  59. Wang, W.; Yan, Y.; Tian, F. Numerical study on hydrodynamics for a non-sinusoidal forced oscillating hydrofoil based on an immersed boundary method. Ocean Eng. 2018, 147, 606–620. [Google Scholar] [CrossRef]
  60. Li, W.; Wang, W.; Yan, Y.; Tian, F. Effects of pitching motion profile on energy harvesting performance of a semi-active flapping foil using immersed boundary method. Ocean Eng. 2018, 163, 94–106. [Google Scholar] [CrossRef]
  61. Xie, F.; Zheng, H.; Deng, J.; Zheng, Y. Vortex induced vibration of a circular cylinder with a filament by using penalty immersed boundary method. Ocean Eng. 2019, 186, 106078. [Google Scholar] [CrossRef]
  62. Ma, Y.; Tucker, P.G. Filtered geometry modelling for fan-intake interaction based on the immersed boundary method. In Proceedings of the 13th European Conference on Turbomachinery Fluid Dynamics & Thermodynamics, Lausanne, Switzerland, 8–12 April 2019. [Google Scholar] [CrossRef]
  63. Wu, X.; Chen, F.; Zhou, S. Stationary and flow-induced vibration of two elliptic cylinders in tandem by immersed boundary-MRT lattice Boltzmann flux solver. J. Fluids Struct. 2019, 91, 102762. [Google Scholar] [CrossRef]
  64. Wang, M.; Avital, E.; Bai, X.; Ji, C.; Xu, D.; Williams, J.; Munjiza, A. Fluid-structure interaction of flexible submerged vegetation stems and kinetic turbine blades. Comput. Part. Mech. 2019, 7, 839–848. [Google Scholar] [CrossRef]
  65. Chen, Y.; Ryu, J.; Liu, Y.; Sung, H.J. Flapping dynamics of vertically clamped three-dimensional flexible flags in a Poiseuille flow. Phys. Fluids 2020, 32, 071905. [Google Scholar] [CrossRef]
  66. Zhao, E.; Sun, J.; Tang, Y.; Mu, L.; Jiang, H. Numerical investigation of tsunami wave impacts on different coastal bridge decks using immersed boundary method. Ocean Eng. 2020, 201, 107132. [Google Scholar] [CrossRef]
  67. Zhang, J.; Sung, H.J.; Huang, W. Specialization of tuna: A numerical study on the function of caudal keels. Phys. Fluids 2020, 32, 111902. [Google Scholar] [CrossRef]
  68. Luo, Y.; Li, X.; Hao, W. Projection-based model reduction for the immersed boundary method. Int. J. Numer. Methods Biomed. Eng. 2021, 38, e3558. [Google Scholar] [CrossRef]
  69. Kasbaoui, M.H.; Kulkarni, T.; Bisetti, F. Direct numerical simulations of the swirling von Kármán flow using a semi-implicit moving immersed boundary method. Comput. Fluids 2021, 230, 105132. [Google Scholar] [CrossRef]
  70. Dong, Z.; Huang, Q.; Pan, G.; Yang, L.; Huang, W. Vortex dynamics and hydrodynamic performance enhancement mechanism in batoid fish oscillatory swimming. J. Fluid Mech. 2021, 930, A28. [Google Scholar] [CrossRef]
  71. Ai, C.; Ma, Y.; Yuan, C.; Dong, G. Non-hydrostatic model for internal wave generations and propagations using immersed boundary method. Ocean. Eng. 2021, 225, 108801. [Google Scholar] [CrossRef]
  72. Yan, H.; Zhang, G.; Wang, S.; Hui, D.; Zhou, B. Simulation of vortex shedding around cylinders by immersed boundary-lattice Boltzmann flux solver. Appl. Ocean Res. 2021, 114, 102763. [Google Scholar] [CrossRef]
  73. Tian, Z.; Zhang, A.; Liu, Y.; Wang, S. Transient fluid-solid interaction with the improved penalty immersed boundary method. Ocean Eng. 2021, 236, 109537. [Google Scholar] [CrossRef]
  74. Zhao, D.; Cui, J.; Dong, L.; Wang, P.; Xiao, Z.; Liu, S.; Chen, L. Drag reduction characteristics of the skin made of micro floating raft arrays based on immersed boundary method. Mech. Based Des. Struct. Mach. 2021, 51, 4833–4846. [Google Scholar] [CrossRef]
  75. Yaswanth, D.; Maniyeri, R. Numerical study of oscillating lid driven cavity with the presence of an obstacle using immersed boundary method. Mater. Today Proc. 2022, 66, 2580–2586. [Google Scholar] [CrossRef]
  76. Mazharmanesh, S.; Young, J.; Tian, F.; Ravi, S.; Lai, J. Energy harvesting of inverted piezoelectric flags in an oscillating flow. J. Fluids Struct. 2022, 115, 103762. [Google Scholar] [CrossRef]
  77. Karimnejad, S.; Delouei, A.A.; He, F. Coupling immersed boundary and lattice Boltzmann method for modelling multi-body interactions subjected to pulsatile flow. Math. Methods Appl. Sci. 2022, 46, 6767–6786. [Google Scholar] [CrossRef]
  78. Mao, Q.; Zhao, J.; Liu, Y.; Sung, H.J. Drag reduction by a flexible hairy coating. J. Fluid Mech. 2022, 946, A5. [Google Scholar] [CrossRef]
  79. Zhang, J.; Sung, H.J.; Huang, W. Hydrodynamic interaction of dorsal fin and caudal fin in swimming tuna. Bioinspir. Biomim. 2022, 17, 066004. [Google Scholar] [CrossRef]
  80. Yu, X.; Yu, M. A volume penalization immersed boundary method for flow interactions with aquatic vegetation. Adv. Water Resour. 2022, 161, 104120. [Google Scholar] [CrossRef]
  81. Jin, Q.; Hudson, D.A.; Temarel, P. A combined volume of fluid and immersed boundary method for free surface simulations induced by solitary waves. Ocean Eng. 2022, 245, 110560. [Google Scholar] [CrossRef]
  82. Fang, D.; Huang, Z.; Zhang, J.; Hu, Z.; Tan, J. Flow pattern investigation of bionic fish by immersed boundary–lattice Boltzmann method and dynamic mode decomposition. Ocean Eng. 2022, 248, 110823. [Google Scholar] [CrossRef]
  83. Huang, Z.; Cheng, Y.; Wu, J.; Diao, W.; Huai, W. FSI simulation of dynamics of fish passing through a tubular turbine based on the immersed boundary-lattice Boltzmann coupling scheme. J. Hydrodyn. 2022, 34, 135–147. [Google Scholar] [CrossRef]
  84. Xiao, Y.; Zhang, G.; Hui, D.; Yan, H.; Feng, S.; Wang, S. Numerical simulation for water entry and exit of rigid bodies based on the immersed boundary-lattice Boltzmann method. J. Fluids Struct. 2022, 109, 103486. [Google Scholar] [CrossRef]
  85. Mi, S.; Wang, M.; Avital, E.; Williams, J.; Chatjigeorgiou, I.K. An implicit Eulerian–Lagrangian model for flow-net interaction using immersed boundary method in OpenFOAM. Ocean Eng. 2022, 264, 112843. [Google Scholar] [CrossRef]
  86. Stival, L.J.L.; Brinkerhoff, J.; Vedovotto, J.M.; De Andrade, F.O. Wake modelling and simulation of an experimental wind turbine using large eddy simulation coupled with immersed boundary method alongside a dynamic adaptive mesh refinement. Energy Convers. Manag. 2022, 268, 115938. [Google Scholar] [CrossRef]
  87. Mazharmanesh, S.; Young, J.; Tian, F.; Ravi, S.; Lai, J. Coupling performance of two tandem and side-by-side inverted piezoelectric flags in an oscillating flow. J. Fluids Struct. 2023, 119, 103874. [Google Scholar] [CrossRef]
  88. Luo, P.; Zhang, J. Numerical simulating wave propagation over solid obstructions using a non-hydrostatic free surface flow model combined immersed boundary method. Ocean Eng. 2023, 269, 113526. [Google Scholar] [CrossRef]
  89. Dong, Z.; Huang, W. Hydrodynamics of a swimming batoid fish at Reynolds numbers up to 148,000. J. Fluid Mech. 2023, 963, A16. [Google Scholar] [CrossRef]
  90. Guo, W.; Hou, G. Combined immersed boundary and discrete unified gas kinetic scheme for the motion of an autonomous underwater vehicle model with slip over a solid-liquid interface. Ocean Eng. 2023, 285, 115322. [Google Scholar] [CrossRef]
  91. Wu, Z.; Guo, L. A new approach to aircraft ditching analysis by coupling free surface lattice Boltzmann and immersed boundary method incorporating surface tension effects. Ocean Eng. 2023, 286, 115559. [Google Scholar] [CrossRef]
  92. Park, S.; Bak, S.; Kim, P.; Seol, Y. A semi-implicit semi-Lagrangian method for simulating immersed boundary motion under high inertia and elasticity. Appl. Math. Comput. 2023, 459, 128269. [Google Scholar] [CrossRef]
  93. Monteiro, L.M.; Mariano, F.P. Flow Modeling over Airfoils and Vertical Axis Wind Turbines Using Fourier Pseudo-Spectral Method and Coupled Immersed Boundary Method. Axioms 2023, 12, 212. [Google Scholar] [CrossRef]
  94. Liu, Z.; Feng, X.; Wang, L.; Tian, F. An Immersed Boundary-Lattice Boltzmann Method Based on the Adaptive Mesh Refinement for Free Surface Flows in Ocean Engineering Applications. Comput. Methods Appl. Mech. Eng. 2023, 392, 114662. [Google Scholar] [CrossRef]
  95. Zhang, H.; Zhao, Y.; Tian, X.; Liu, H. Numerical simulations of the flow past a perforated plate enclosed by a filament using the Immersed Boundary Method. Ocean Eng. 2024, 304, 117860. [Google Scholar] [CrossRef]
  96. Mao, Q.; Liu, Y.; Sung, H.J. Snap-through dynamics of a buckled flexible filament in a uniform flow. J. Fluid Mech. 2023, 969, A33. [Google Scholar] [CrossRef]
  97. Borazjani, I.; Sotiropoulos, F. Vortex-induced vibrations of two cylinders in tandem arrangement in the proximity–wake interference region. J. Fluid Mech. 2009, 621, 321–364. [Google Scholar] [CrossRef]
  98. Seo, J.; Mittal, R. A sharp-interface immersed boundary method with improved mass conservation and reduced spurious pressure oscillations. J. Comput. Phys. 2011, 230, 7347–7363. [Google Scholar] [CrossRef] [PubMed]
  99. Griffith, M.D.; Lo Jacono, D.; Sheridan, J.; Leontini, J. Passive heaving of elliptical cylinders with active pitching—From cylinders towards flapping foils. J. Fluids Struct. 2016, 67, 124–141. [Google Scholar] [CrossRef]
  100. Griffith, M.D.; Leontini, J. Sharp interface immersed boundary methods and their application to vortex-induced vibration of a cylinder. J. Fluids Struct. 2017, 72, 38–58. [Google Scholar] [CrossRef]
  101. Khalili, M.; Larsson, M.; Müller, B. Immersed boundary method for viscous compressible flows around moving bodies. Comput. Fluids 2018, 170, 77–92. [Google Scholar] [CrossRef]
  102. Mishra, R.; Kulkarni, S.S.; Bhardwaj, R.; Thompson, M.C. Response of a linear viscoelastic splitter plate attached to a cylinder in laminar flow. J. Fluids Struct. 2019, 87, 284–301. [Google Scholar] [CrossRef]
  103. Majumdar, D.; Bose, C.; Sarkar, S. Capturing the dynamical transitions in the flow-field of a flapping foil using Immersed Boundary Method. J. Fluids Struct. 2020, 95, 102999. [Google Scholar] [CrossRef]
  104. Narváez, G.; Schettini, E.B.C.; Silvestrini, J.H. Numerical simulation of flow-induced vibration of two cylinders elastically mounted in tandem by immersed moving boundary method. Appl. Math. Model. 2020, 77, 1331–1347. [Google Scholar] [CrossRef]
  105. Xu, Y.; Bingham, H.B.; Shao, Y. Finite difference solutions for nonlinear water waves using an immersed boundary method. Int. J. Numer. Methods Fluids 2020, 93, 1143–1162. [Google Scholar] [CrossRef]
  106. Xin, J.; Chen, Z.; Fan, S.; Shi, F.; Jin, Q. Numerical simulation of nonlinear sloshing in a prismatic tank by a Cartesian grid based three-dimensional multiphase flow model. Ocean Eng. 2020, 213, 107629. [Google Scholar] [CrossRef]
  107. Kundu, A.; Soti, A.K.; Garg, H.; Bhardwaj, R.; Thompson, M.C. Computational modelling and analysis of flow-induced vibration of an elastic splitter plate using a sharp-interface immersed boundary method. SN Appl. Sci. 2020, 2, 1110. [Google Scholar] [CrossRef]
  108. Kwon, H.; Lee, H.; Chang, S. High-order WENO Schemes with an Immersed Boundary Method for Shallow Water Equations on the Tsunami Mitigation with Configurations of Cylinder Array. KSCE J. Civ. Eng. 2020, 25, 1–11. [Google Scholar] [CrossRef]
  109. Tsai, Y.; Lo, D.C. A Ghost-Cell immersed boundary method for Wave–Structure interaction using a Two-Phase flow model. Water 2020, 12, 3346. [Google Scholar] [CrossRef]
  110. Seshadri, P.K.; De, A. A robust sharp interface based immersed boundary framework for moving body problems with applications to laminar incompressible flows. Comput. Math. Appl. 2021, 83, 24–56. [Google Scholar] [CrossRef]
  111. Robaux, F.; Benoît, M. Development and validation of a numerical wave tank based on the Harmonic Polynomial Cell and Immersed Boundary methods to model nonlinear wave-structure interaction. J. Comput. Phys. 2021, 446, 110560. [Google Scholar] [CrossRef]
  112. Badhurshah, R.; Bhardwaj, R.; Bhattacharya, A. Numerical simulation of Vortex-Induced Vibration with bistable springs: Consistency with the Equilibrium Constraint. J. Fluids Struct. 2021, 103, 103280. [Google Scholar] [CrossRef]
  113. Tong, C.W.; Shao, Y.; Bingham, H.B.; Hanssen, F.C.W. An adaptive harmonic polynomial cell method with immersed boundaries: Accuracy, stability, and applications. Int. J. Numer. Methods Eng. 2021, 122, 2945–2980. [Google Scholar] [CrossRef]
  114. Hanssen, F.W.; Greco, M. A potential flow method combining immersed boundaries and overlapping grids: Formulation, validation and verification. Ocean Eng. 2021, 227, 108841. [Google Scholar] [CrossRef]
  115. Sharma, G.; Garg, H.; Bhardwaj, R. Flow-induced vibrations of elastically-mounted C- and D-section cylinders. J. Fluids Struct. 2022, 109, 103501. [Google Scholar] [CrossRef]
  116. Giannenas, A.E.; Laizet, S.; Rigas, G. Harmonic forcing of a laminar bluff body wake with rear pitching flaps. J. Fluid Mech. 2022, 945, A5. [Google Scholar] [CrossRef]
  117. Khedkar, K.; Bhalla, A.P.S. A model predictive control (MPC)-integrated multiphase immersed boundary (IB) framework for simulating wave energy converters (WECs). Ocean. Eng. 2022, 260, 111908. [Google Scholar] [CrossRef]
  118. Song, Y.; Xu, Y.; Ismail, H.; Liu, X. Scour modelling based on immersed boundary method: A pathway to the practical use of three-dimensional scour models. Coast. Eng. 2022, 171, 104037. [Google Scholar] [CrossRef]
  119. Gómez, H.; Narváez, G.; Schettini, E.B.C. Vortex induced vibration of four cylinders configurations at critical spacing in 0° and 45° flow incidence angle. Ocean Eng. 2022, 252, 111134. [Google Scholar] [CrossRef]
  120. Badhurshah, R.; Bhardwaj, R.; Bhattacharya, A. Energy extraction via Vortex-Induced Vibrations: The effect of spring bistability. J. Fluids Struct. 2022, 114, 103708. [Google Scholar] [CrossRef]
  121. Ji, R.; Sun, K.; Zhang, J.; Zhu, R.; Wang, S. A novel actuator line-immersed boundary (AL-IB) hybrid approach for wake characteristics prediction of a horizontal-axis wind turbine. Energy Convers. Manag. 2022, 253, 115193. [Google Scholar] [CrossRef]
  122. Xu, Y.; Bingham, H.B.; Shao, Y. A high-order finite difference method with immersed-boundary treatment for fully-nonlinear wave–structure interaction. Appl. Ocean Res. 2023, 134, 103535. [Google Scholar] [CrossRef]
  123. Pandey, A.K.; Sharma, G.; Bhardwaj, R. Flow-induced reconfiguration and cross-flow vibrations of an elastic plate and implications to energy harvesting. J. Fluids Struct. 2023, 122, 103977. [Google Scholar] [CrossRef]
  124. Kanchan, M.; Lewis, D.; Varma, A. Fluid-induced deformation of elastic plate mounted vertically in viscous flow and oscillation frequency prediction using artificial neural networks. Ocean Eng. 2024, 296, 116920. [Google Scholar] [CrossRef]
  125. Chern, M.; Wang, C.; Wei, Z.; Lu, P. Numerical Investigation of a Pitching NACA 0012 Wing with Plasma-Based Flow Control Using Prediction–Correction Direct-Forcing Immersed Boundary Method. J. Aerosp. Eng. 2023, 36, 1032. [Google Scholar] [CrossRef]
  126. Yang, B.; Cheng, Q.; Song, M. The study of vertical-axis wind turbine based on immersed boundary method. Energy Explor. Exploit. 2023, 42, 692–711. [Google Scholar] [CrossRef]
  127. Yu, X.; Shao, Y.; Fuhrman, D.R.; Zhang, Y. A viscous numerical wave tank based on immersed-boundary generalized harmonic polynomial cell (IB-GHPC) method: Accuracy, validation and application. Coast. Eng. 2023, 180, 104273. [Google Scholar] [CrossRef]
  128. Sundar, R.; Majumdar, D.; Lucor, D.; Sarkar, S. Physics-informed neural networks modelling for systems with moving immersed boundaries: Application to an unsteady flow past a plunging foil. arXiv 2023, arXiv:2306.13395. [Google Scholar] [CrossRef]
  129. Kolahdouz, E.M.; Wells, D.; Rossi, S.; Aycock, K.I.; Craven, B.A.; Griffith, B.E. A sharp interface Lagrangian-Eulerian method for flexible-body fluid-structure interaction. J. Comput. Phys. 2023, 488, 112174. [Google Scholar] [CrossRef] [PubMed]
  130. Zargaran, A.; Dolshanskiy, W.; Stepanyuk, A.; Pauer, W.; Janoske, U. A hybrid approach based on Lagrangian particles and immersed-boundary method to characterize rotor–stator mixing systems for high viscous mixtures. Chem. Eng. J. 2023, 473, 145062. [Google Scholar] [CrossRef]
  131. Li, H.; Feng, J.; Zheng, Y.; Xu, H.; Chen, H.; Binama, M.; Kan, K. A computational method for complex-shaped hydraulic turbomachinery flow based on the immersed boundary method. AIP Adv. 2023, 13, 085121. [Google Scholar] [CrossRef]
  132. Joachim, J.; Daunais, C.; Bibeau, V.; Heltai, L.; Blais, B. A parallel and adaptative Nitsche immersed boundary method to simulate viscous mixing. J. Comput. Phys. 2023, 488, 112189. [Google Scholar] [CrossRef]
  133. Kou, J.; Ferrer, E. A combined volume penalisation/selective frequency damping approach for immersed boundary methods: Application to moving geometries. J. Comput. Phys. 2023, 472, 111678. [Google Scholar] [CrossRef]
  134. Agrawal, V.; Kulachenko, A.; Scapin, N.; Tammisola, O.; Brandt, L. An efficient isogeometric/finite-difference immersed boundary method for the fluid–structure interactions of slender flexible structures. Comput. Methods Appl. Mech. Eng. 2024, 418, 116495. [Google Scholar] [CrossRef]
  135. Wang, L.; Tian, F. Numerical study of sound generation by three-dimensional flexible flapping wings during hovering flight. J. Fluids Struct. 2020, 99, 103165. [Google Scholar] [CrossRef]
  136. Wang, L.; Tian, F. Numerical study of flexible flapping wings with an immersed boundary method: Fluid–structure–acoustics interaction. J. Fluids Struct. 2019, 90, 396–409. [Google Scholar] [CrossRef]
  137. Pinelli, A.; Omidyeganeh, M.; Brücker, C.; Revell, A.; Sarkar, A.; Alinovi, E. The pelskin project: Part IV—Control of bluff body wakes using hairy filaments. Meccanica 2017, 52, 1503–1514. [Google Scholar] [CrossRef]
  138. Ren, W.; Chang, S.; Yang, W. An efficient immersed boundary method for thermal flow problems with heat flux boundary conditions. Int. J. Heat Mass Transf. 2013, 64, 694–705. [Google Scholar] [CrossRef]
  139. Mazharmanesh, S.; Young, J.; Tian, F.; Lai, J. Energy harvesting of two inverted piezoelectric flags in tandem, side-by-side and staggered arrangements. Int. J. Heat Fluid Flow 2020, 83, 108589. [Google Scholar] [CrossRef]
  140. Tong, Z.; Sun, Q.; Dong, L.; Gu, Z. Two-dimensional numerical model for predicting fouling shape growth based on immersed boundary method and lattice Boltzmann method. Appl. Therm. Eng. 2020, 179, 115755. [Google Scholar] [CrossRef]
  141. Abaszadeh, M.; Ali, S.; Delouei, A.A.; Amiri, H. Analysis of radiative heat transfer in two-dimensional irregular geometries by developed immersed boundary–lattice Boltzmann method. J. Quant. Spectrosc. Radiat. Transf. 2022, 280, 108086. [Google Scholar] [CrossRef]
  142. Jiang, M.; Li, J.; Liu, Z. A simple and efficient parallel immersed boundary-lattice Boltzmann method for fully resolved simulations of incompressible settling suspensions. Comput. Fluids 2022, 237, 105322. [Google Scholar] [CrossRef]
  143. Tao, S.; Wang, L.; He, Q.; Chen, J.; Luo, J. Lattice Boltzmann simulation of complex thermal flows via a simplified immersed boundary method. J. Comput. Sci. 2022, 65, 101878. [Google Scholar] [CrossRef]
  144. Haeri, S.; Shrimpton, J.S. A new implicit fictitious domain method for the simulation of flow in complex geometries with heat transfer. J. Comput. Phys. 2013, 237, 21–45. [Google Scholar] [CrossRef]
  145. Chen, Z.; Shu, C.; Yang, L.M.; Zhao, X.; Liu, N.Y. Immersed boundary–simplified thermal lattice BoltzFsundarnn method for incompressible thermal flows. Phys. Fluids 2020, 32, 013605. [Google Scholar] [CrossRef]
  146. Xu, T.; Choi, J.I. Efficient monolithic immersed boundary projection method for incompressible flows with heat transfer. J. Comput. Phys. 2023, 477, 111929. [Google Scholar] [CrossRef]
  147. Hosseini, S.; Aghebatandish, S.; Dadvand, A.; Khoo, B.C. An immersed boundary-lattice Boltzmann method with multi relaxation time for solving flow-induced vibrations of an elastic vortex generator and its effect on heat transfer and mixing. Chem. Eng. J. 2021, 405, 126652. [Google Scholar] [CrossRef]
  148. Wu, B.; Lu, J.; Lee, H.C.; Shu, C.; Wan, M. An explicit boundary condition-enforced immersed boundary-reconstructed thermal lattice Boltzmann flux solver for thermal–fluid–structure interaction problems with heat flux boundary conditions. J. Comput. Phys. 2023, 485, 112106. [Google Scholar] [CrossRef]
  149. Chen, Y.; Yang, J.; Liu, Y.; Sung, H.J. Heat transfer enhancement in a poiseuille channel flow by using multiple wall-mounted flexible flags. Int. J. Heat Mass Transf. 2020, 163, 120447. [Google Scholar] [CrossRef]
  150. Wang, L.; Tian, F.; Young, J.B. An immersed boundary method for the fluid--structure--thermal interaction in rarefied gas flow. arXiv 2023, arXiv:2305.11454. [Google Scholar] [CrossRef]
  151. Zhang, H.; Yuan, H.; Trias, F.X.; Yu, A.; Tan, Y.; Oliva, A. Particulate Immersed Boundary Method for complex fluid–particle interaction problems with heat transfer. Comput. Math. Appl. 2016, 71, 391–407. [Google Scholar] [CrossRef]
  152. Wu, B.; Chen, S.; Wan, M. An implicit immersed boundary method for Robin boundary condition. Int. J. Mech. Sci. 2024, 261, 108694. [Google Scholar] [CrossRef]
  153. Pacheco, J.R.; Pacheco-Vega, A.; Rodić, T.; Peck, R. Numerical simulations of heat transfer and fluid flow problems using an Immersed-Boundary Finite-Volume method on NonStaggered grids. Numer. Heat Transf. Part B: Fundam. 2005, 48, 1–24. [Google Scholar] [CrossRef]
  154. Soti, A.K.; Bhardwaj, R.; Sheridan, J. Flow-induced deformation of a flexible thin structure as manifestation of heat transfer enhancement. Int. J. Heat Mass Transf. 2015, 84, 1070–1081. [Google Scholar] [CrossRef]
  155. Garg, H.; Soti, A.K.; Bhardwaj, R. A sharp interface immersed boundary method for vortex-induced vibration in the presence of thermal buoyancy. Phys. Fluids 2018, 30, 023603. [Google Scholar] [CrossRef]
  156. Garg, H.; Soti, A.K.; Bhardwaj, R. Thermal buoyancy induced suppression of wake-induced vibration. Int. Commun. Heat Mass Transf. 2020, 118, 104790. [Google Scholar] [CrossRef]
  157. Lou, J.; Johnston, J.; Tilton, N. Application of projection and immersed boundary methods to simulating heat and mass transport in membrane distillation. Comput. Fluids 2020, 212, 104711. [Google Scholar] [CrossRef]
  158. Mohammadi, M.; Nassab, S.A.G. Application of the immersed boundary method in solution of radiative heat transfer problems. J. Quant. Spectrosc. Radiat. Transf. 2021, 260, 107467. [Google Scholar] [CrossRef]
  159. Mohammadi, M.; Nassab, S.A.G. Solution of radiative-convective heat transfer in irregular geometries using hybrid lattice Boltzmann-finite volume and immersed boundary method. Int. Commun. Heat Mass Transf. 2021, 128, 105595. [Google Scholar] [CrossRef]
  160. Riahi, H.; Goncalvès, E.; Meldi, M. A Discrete Immersed Boundary Method for the numerical simulation of heat transfer in compressible flows. arXiv 2023, arXiv:2301.09349. [Google Scholar] [CrossRef]
  161. Ahn, J.; Song, J.; Lee, J.S. An immersed boundary method for conjugate heat transfer involving Melting/Solidification. Int. J. Aeronaut. Space Sci. 2023, 24, 1032–1041. [Google Scholar] [CrossRef]
  162. Cruz, R.V.; Lamballais, É. A versatile immersed boundary method for high-fidelity simulation of Conjugate Heat Transfer. J. Comput. Phys. 2023, 488, 112182. [Google Scholar] [CrossRef]
  163. Wang, D.; Jin, T.; Luo, K.; Fan, J. An improved direct-forcing immersed boundary method for simulations of flow and heat transfer in particle-laden flows. Int. J. Multiph. Flow 2022, 153, 104139. [Google Scholar] [CrossRef]
  164. Narváez, G.; Lamballais, É.; Schettini, E.B.C. Simulation of turbulent flow subjected to conjugate heat transfer via a dual immersed boundary method. Comput. Fluids 2021, 229, 105101. [Google Scholar] [CrossRef]
  165. Wu, B.; Lu, J.; Lee, H.C.; Shu, C.; Wan, M. An explicit immersed boundary-reconstructed thermal lattice Boltzmann flux solver for thermal–fluid-structure interaction problems. Int. J. Mech. Sci. 2022, 235, 107704. [Google Scholar] [CrossRef]
  166. Zhao, Z.; Yan, J. Enriched immersed boundary method (EIBM) for interface-coupled multi-physics and applications to convective conjugate heat transfer. Comput. Methods Appl. Mech. Eng. 2022, 401, 115667. [Google Scholar] [CrossRef]
  167. Ou, Z.; Chi, C.; Guo, L.; Thévenin, D. A directional ghost-cell immersed boundary method for low Mach number reacting flows with interphase heat and mass transfer. J. Comput. Phys. 2022, 468, 111447. [Google Scholar] [CrossRef]
  168. Xia, J.Y.; Luo, K.; Fan, J. A ghost-cell based high-order immersed boundary method for inter-phase heat transfer simulation. Int. J. Heat Mass Transf. 2014, 75, 302–312. [Google Scholar] [CrossRef]
  169. Tao, S.; Wang, L.; He, Q.; Chen, J.; Luo, J. A sharp interface immersed boundary-discrete unified gas kinetic scheme for fluid-solid flows with heat transfer. Int. Commun. Heat Mass Transf. 2022, 139, 106424. [Google Scholar] [CrossRef]
  170. Fernandez, D.; Husain, S.Z.; Floryan, J.M. Immersed boundary conditions method for heat conduction problems in slots with time-dependent geometry. Int. J. Numer. Methods Fluids 2011, 67, 478–500. [Google Scholar] [CrossRef]
  171. Shrivastava, M.; Agrawal, A.; Sharma, A. A novel level Set-Based Immersed-Boundary method for CFD simulation of Moving-Boundary problems. Numer. Heat Transf. Part B Fundam. 2013, 63, 304–326. [Google Scholar] [CrossRef]
  172. Ménez, L.; Parnaudeau, P.; Béringhier, M.; Da Silva, E.G. Assessment of volume penalization and immersed boundary methods for compressible flows with various thermal boundary conditions. J. Comput. Phys. 2023, 493, 112465. [Google Scholar] [CrossRef]
  173. Zhang, Y.; Pan, G.; Zhang, Y.; Haeri, S. A multi-physics peridynamics-DEM-IB-CLBM framework for the prediction of erosive impact of solid particles in viscous fluids. Comput. Methods Appl. Mech. Eng. 2019, 352, 675–690. [Google Scholar] [CrossRef]
  174. Yang, S.; Xu, C.; Huang, W.; Wang, L. Transient growth in turbulent particle-laden channel flow. Acta Mech. Sin. 2019, 36, 1–11. [Google Scholar] [CrossRef]
  175. Wang, M.; Feng, Y.T.; Qu, T.; Zhao, T. A coupled polygonal DEM-LBM technique based on an immersed boundary method and energy-conserving contact algorithm. Powder Technol. 2021, 381, 101–109. [Google Scholar] [CrossRef]
  176. Zhang, C.; Zhang, H.; Zhao, Y.; Yang, C. An immersed boundary-lattice Boltzmann model for simulation of deposited particle patterns in an evaporating sessile droplet with dispersed particles. Int. J. Heat Mass Transf. 2021, 181, 121905. [Google Scholar] [CrossRef]
  177. Romanus, R.S.; Lugarini, A.; Franco, A. An immersed boundary-lattice Boltzmann framework for fully resolved simulations of non-spherical particle settling in unbounded domain. Comput. Math. Appl. 2021, 102, 206–219. [Google Scholar] [CrossRef]
  178. Wang, W.; Wang, J.; Cui, G.; Pei, J.; Yan, Y. A numerical study on elliptical particle deposition with an immersed boundary-lattice Boltzmann method. Comput. Fluids 2022, 246, 105644. [Google Scholar] [CrossRef]
  179. Fukui, T.; Kawaguchi, M. Numerical study of microscopic particle arrangement of suspension flow in a narrow channel for the estimation of macroscopic rheological properties. Adv. Powder Technol. 2022, 33, 103855. [Google Scholar] [CrossRef]
  180. Cheng, Z.; Wachs, A. An immersed boundary/multi-relaxation time lattice Boltzmann method on adaptive octree grids for the particle-resolved simulation of particle-laden flows. J. Comput. Phys. 2022, 471, 111669. [Google Scholar] [CrossRef]
  181. Yadav, P.; Ghosh, S. Numerical Studies of settling of an impermeable and permeable planktonic particle using Immersed boundary method (IBM). Eur. Phys. J. Plus 2022, 137, 740. [Google Scholar] [CrossRef]
  182. Kawaguchi, M.; Fukui, T.; Morinishi, K. Comparative study of the virtual flux method and immersed boundary method coupled with regularized lattice Boltzmann method for suspension flow simulations. Comput. Fluids 2022, 246, 105615. [Google Scholar] [CrossRef]
  183. Ghosh, S.; Panghal, R. Study of gravitational settling of a flexible circular structure using immersed boundary method. Comput. Appl. Math. 2022, 41, 339. [Google Scholar] [CrossRef]
  184. Chang, H.; Chen, A.; Ge, B. A hybrid method of peridynamic differential operator-based Eulerian particle method–immersed boundary method for fluid–structure interaction. Comput. Part. Mech. 2023, 10, 1309–1322. [Google Scholar] [CrossRef]
  185. Panghal, R.; Ghosh, S. Study of gravitational sedimentation of flexible, permeable circular, and planktonic particles applying the immersed boundary method. Int. J. Sediment Res. 2023, 38, 643–652. [Google Scholar] [CrossRef]
  186. Zhang, C.; Yin, S.; Zhang, H.; Yang, C. Simulation of a sessile nanofluid droplet freezing with an immersed boundary-lattice Boltzmann model. Int. J. Multiph. Flow 2023, 167, 104553. [Google Scholar] [CrossRef]
  187. Yadav, P.; Ghosh, S.; Panghal, R. Numerical studies of settling of a permeable particle of semi-torus shape applying immersed boundary method (IBM). AIP Conf. Proc. 2023, 2872, 120075. [Google Scholar] [CrossRef]
  188. Ghosh, A.; Gabbana, A.; Null, H.W.; Toschi, F. Effective force stabilising technique for the immersed boundary method. Commun. Comput. Phys. 2023, 33, 349–366. [Google Scholar] [CrossRef]
  189. Patel, J.K.; Natarajan, G. Diffuse interface immersed boundary method for multi-fluid flows with arbitrarily moving rigid bodies. J. Comput. Phys. 2018, 360, 202–228. [Google Scholar] [CrossRef]
  190. Souza, P.R.C.; Neto, H.R.; Villar, M.M.; Vedovotto, J.M.; Cavalini, A.A.; Neto, A.S. Multi-phase fluid–structure interaction using adaptive mesh refinement and immersed boundary method. J. Braz. Soc. Mech. Sci. Eng. 2022, 44, 152. [Google Scholar] [CrossRef]
  191. Niu, X.; Zhou, J.; Xiao, H.; Wang, Y.; Khan, A.; Chen, M.; Li, D.; Yamaguchi, H. A simple diffuse interface immersed-boundary scheme for multiphase flows with curved boundaries. Int. J. Multiph. Flow 2022, 157, 104266. [Google Scholar] [CrossRef]
  192. Wang, Q.; Pan, M.; Tseng, Y.; He, D. An Energy Stable Immersed Boundary Method for Deformable Membrane Problem with Non-uniform Density and Viscosity. J. Sci. Comput. 2023, 94, 30. [Google Scholar] [CrossRef]
  193. Ye, H.; Chen, Y.; Maki, K.J. A Discrete-Forcing immersed boundary method for moving bodies in Air–Water Two-Phase flows. J. Mar. Sci. Eng. 2020, 8, 809. [Google Scholar] [CrossRef]
  194. Bilbao, S. Modeling impedance boundary conditions and acoustic barriers using the immersed boundary method: The one-dimensional case. J. Acoust. Soc. Am. 2023, 153, 2023–2036. [Google Scholar] [CrossRef] [PubMed]
  195. Bilbao, S. Modeling impedance boundary conditions and acoustic barriers using the immersed boundary method: The three-dimensional case. J. Acoust. Soc. Am. 2023, 154, 874–885. [Google Scholar] [CrossRef]
  196. Hou, J.; Zheng, Z.C.; Allen, J.S. Time-domain immersed-boundary simulation of acoustic propagation between two spherical gas bubbles. JASA Express Lett. 2023, 3, 094002. [Google Scholar] [CrossRef]
  197. Nangia, N.; Patankar, N.A.; Bhalla, A.P.S. A DLM immersed boundary method based wave-structure interaction solver for high density ratio multiphase flows. J. Comput. Phys. 2019, 398, 108804. [Google Scholar] [CrossRef]
  198. Cheng, L.; Du, L.; Wang, X.; Sun, X.; Tucker, P.G. A semi-implicit immersed boundary method for simulating viscous flow-induced sound with moving boundaries. Comput. Methods Appl. Mech. Eng. 2021, 373, 113438. [Google Scholar] [CrossRef]
  199. Zeng, Y.; Bhalla, A.P.S.; Shen, L. A subcycling/non-subcycling time advancement scheme-based DLM immersed boundary method framework for solving single and multiphase fluid–structure interaction problems on dynamically adaptive grids. Comput. Fluids 2022, 238, 105358. [Google Scholar] [CrossRef]
  200. Yan, H.; Zhang, G.; Xiao, Y.; Hui, D.; Wang, S. A surface flux correction-based immersed boundary-multiphase lattice Boltzmann flux solver applied to multiphase fluids–structure interaction. Comput. Methods Appl. Mech. Eng. 2022, 400, 115481. [Google Scholar] [CrossRef]
  201. Hori, N.; Rosti, M.E.; Takagi, S. An Eulerian-based immersed boundary method for particle suspensions with implicit lubrication model. Comput. Fluids 2022, 236, 105278. [Google Scholar] [CrossRef]
  202. Liu, X.; Ji, C.; Xu, X.; Xu, D.; Williams, J. Distribution characteristics of inertial sediment particles in the turbulent boundary layer of an open channel flow determined using Voronoï analysis. Int. J. Sediment Res. 2017, 32, 401–409. [Google Scholar] [CrossRef]
  203. Wang, L.; Currao, G.M.D.; Feng, H.; Neely, A.J.; Young, J.; Tian, F. An immersed boundary method for fluid–structure interaction with compressible multiphase flows. J. Comput. Phys. 2017, 346, 131–151. [Google Scholar] [CrossRef]
  204. Cheng, L.; Du, L.; Jing, X.; Sun, X.; Hu, G. A Compressible Immersed Boundary Method For The Flow-Induced Noise Simulation. In Proceedings of the 24th International Congress on Sound and Vibration (ICSV24), London, UK, 23–27 July 2017; p. 423. [Google Scholar]
  205. Zhang, C.; Zhang, W.; Lin, N.; Tang, Y.; Zhao, C.; Gu, J.; Wang, L.; Chen, X.; Qiu, A. A two-phase flow model coupling with volume of fluid and immersed boundary methods for free surface and moving structure problems. Ocean Eng. 2013, 74, 107–124. [Google Scholar] [CrossRef]
  206. Bürchner, T.; Kopp, P.; Kollmannsberger, S.; Rank, E. Immersed boundary parametrizations for full waveform inversion. Comput. Methods Appl. Mech. Eng. 2023, 406, 115893. [Google Scholar] [CrossRef]
  207. Fazli, M.; Rudman, M.; Kuang, S.; Chryss, A. Application of immersed boundary methods to non-Newtonian yield-pseudoplastic flows. Appl. Math. Model. 2023, 124, 532–552. [Google Scholar] [CrossRef]
  208. Chéron, V.; Evrard, F.; Van Wachem, B. A hybrid immersed boundary method for dense particle-laden flows. Comput. Fluids 2023, 259, 105892. [Google Scholar] [CrossRef]
  209. Ido, Y.; Sumiyoshi, H.; Tsutsumi, H. Simulations of behavior of magnetic particles in magnetic functional fluids using a hybrid method of lattice Boltzmann method, immersed boundary method and discrete particle method. Comput. Fluids 2017, 142, 86–95. [Google Scholar] [CrossRef]
  210. Fukui, T.; Kawaguchi, M.; Morinishi, K. A two-way coupling scheme to model the effects of particle rotation on the rheological properties of a semidilute suspension. Comput. Fluids 2018, 173, 6–16. [Google Scholar] [CrossRef]
  211. Barbeau, L.; Étienne, S.; Béguin, C.; Blais, B. Development of a high-order continuous Galerkin sharp-interface immersed boundary method and its application to incompressible flow problems. Comput. Fluids 2022, 239, 105415. [Google Scholar] [CrossRef]
  212. Wu, G.; Chen, S. Simulating spray coating processes by a three-dimensional lattice Boltzmann method-immersed boundary method approach. Chem. Eng. Sci. 2022, 263, 118091. [Google Scholar] [CrossRef]
  213. Duprez, M.; Lleras, V.; Lozinski, A. ϕ-FEM: An optimally convergent and easily implementable immersed boundary method for particulate flows and Stokes equations. ESAIM Math. Model. Numer. Anal. 2023, 57, 1111–1142. [Google Scholar] [CrossRef]
  214. Farooq, H.; Akhtar, I.; Hemmati, A.; Khalid, M.S.U. An accurate immersed boundary method using radial-basis functions for incompressible flows. J. Comput. Phys. 2025, 531, 113928. [Google Scholar] [CrossRef]
  215. Zhao, C.; Yang, Y.; Zhang, T.; Dong, H.; Hou, G. A sharp interface immersed boundary method for flow-induced noise prediction using acoustic perturbation equations. Comput. Fluids 2021, 227, 105032. [Google Scholar] [CrossRef]
  216. Gorges, C.; Brömmer, M.; Velten, C.; Wirtz, S.; Illana, E.; Scherer, V.; Zähringer, K.; Van Wachem, B. Comparing two IBM implementations for the simulation of uniform packed beds. Particuology 2024, 86, 1–12. [Google Scholar] [CrossRef]
  217. Zhou, K.; Balachandar, S. An analysis of the spatio-temporal resolution of the immersed boundary method with direct forcing. J. Comput. Phys. 2021, 424, 109862. [Google Scholar] [CrossRef]
  218. Xie, F.; Qu, Y.; Islam, A.; Meng, G. A sharp-interface Cartesian grid method for time-domain acoustic scattering from complex geometries. Comput. Fluids 2020, 202, 104498. [Google Scholar] [CrossRef]
  219. Zhao, C.; Zhang, T.; Hou, G. Finite-difference time-domain modelling for underwater acoustic scattering applications based on immersed boundary method. Appl. Acoust. 2022, 193, 108764. [Google Scholar] [CrossRef]
  220. Isoz, M.; Šourek, M.K.; Studeník, O.; Kočí, P. Hybrid fictitious domain-immersed boundary solver coupled with discrete element method for simulations of flows laden with arbitrarily-shaped particles. Comput. Fluids 2022, 244, 105538. [Google Scholar] [CrossRef]
  221. Qin, J.; Null, X.Y.; Li, Z. Hybrid diffuse and sharp interface immersed boundary methods for particulate flows in the presence of complex boundaries. Commun. Comput. Phys. 2022, 31, 1242–1271. [Google Scholar] [CrossRef]
  222. Giahi, M.; Bergstrom, D.E. A critical assessment of the immersed boundary method for modeling flow around fixed and moving bodies. Comput. Fluids 2023, 256, 105841. [Google Scholar] [CrossRef]
  223. Zeng, J.; Li, H.; Zhang, D. Direct numerical simulation of proppant transport in hydraulic fractures with the immersed boundary method and multi-sphere modeling. Appl. Math. Model. 2021, 91, 590–613. [Google Scholar] [CrossRef]
  224. Caunt, E.; Nelson, R.; Luporini, F.; Gorman, G.J. A Novel Immersed Boundary Approach for Irregular Topography with Acoustic Wave Equations. arXiv 2023, arXiv:2309.03600. [Google Scholar] [CrossRef]
  225. Uhlmann, M. An immersed boundary method with direct forcing for the simulation of particulate flows. J. Comput. Phys. 2005, 209, 448–476. [Google Scholar] [CrossRef]
  226. Mentzoni, F.; Kristiansen, T. Two-dimensional experimental and numerical investigations of perforated plates in oscillating flow, orbital flow and incident waves. Appl. Ocean Res. 2020, 97, 102078. [Google Scholar] [CrossRef]
  227. Haeri, S.; Shrimpton, J.S. On the application of immersed boundary, fictitious domain and body-conformal mesh methods to many particle multiphase flows. Int. J. Multiph. Flow 2012, 40, 38–55. [Google Scholar] [CrossRef]
  228. Farea, A.; Khan, S.; Daryani, R.; Ersan, E.C.; Celebi, M.S. Learning Fluid-Structure Interaction Dynamics with Physics-Informed Neural Networks and Immersed Boundary Methods. arXiv 2025, arXiv:2505.18565. [Google Scholar] [CrossRef]
  229. Zuber, D.; Larsson, J.; Brehm, C.; McQuaid, J.; Van Noordt, W.; Min, B.; Wake, B.E. Wall-Modeled Large Eddy simulation using the immersed boundary method of the HVAB rotor. In Proceedings of the AIAA SCITECH 2022 Forum, Orlando, FL, USA, 6–10 January 2025. [Google Scholar] [CrossRef]
  230. Gopakumar, V.; Gray, A.; Zanisi, L.; Nunn, T.; Pamela, S.; Giles, D.; Kusner, M.J.; Deisenroth, M.P. Calibrated Physics-Informed uncertainty quantification. arXiv 2025, arXiv:2502.04406. [Google Scholar] [CrossRef]
  231. Zhou, Z.; Zhang, F.; Yang, X. A features-embedded-learning immersed boundary model for large-eddy simulation of turbulent flows with complex boundaries. arXiv 2025, arXiv:2506.21985. [Google Scholar] [CrossRef]
  232. Powar, O.; Arun Pa, H.; Kumar, A.M.; Kanchan, M.; Karthik, B.M.; Mangalore, P.; Santhya, M. Recent developments in the immersed boundary Method for Complex Fluid–Structure Interactions: A review. Fluids 2025, 10, 134. [Google Scholar] [CrossRef]
Figure 1. PRISMA-style flowchart illustrating article selection process for the review.
Figure 1. PRISMA-style flowchart illustrating article selection process for the review.
Fluids 10 00217 g001
Figure 2. (a) Diffuse-interface methods and (b) Sharp-interface methods along with respective boundary interfaces.
Figure 2. (a) Diffuse-interface methods and (b) Sharp-interface methods along with respective boundary interfaces.
Fluids 10 00217 g002
Figure 3. Schematic illustration of a flexible filament placed freely in the channel mid-plane, subjected to viscous shear flow. The filament is initially inclined at an angle of 5° (α) to induce asymmetry in its motion. Non-dimensional parameters include filament length (L), shear rate (K), and bending rigidity (Kb) (Kanchan and Maniyeri, 2019) [9].
Figure 3. Schematic illustration of a flexible filament placed freely in the channel mid-plane, subjected to viscous shear flow. The filament is initially inclined at an angle of 5° (α) to induce asymmetry in its motion. Non-dimensional parameters include filament length (L), shear rate (K), and bending rigidity (Kb) (Kanchan and Maniyeri, 2019) [9].
Fluids 10 00217 g003
Figure 4. Deformation of a filament at different time points for various bending rigidity values (Kb), illustrating distinct deformation modes: rigid (Kb = 1.0), springy (Kb = 0.7), C-shape (Kb = 2.5 × 10−2), S-shape (Kb = 6 × 10−3), and complex deformations (Kb = 5 × 10−4). As bending rigidity decreases, the filament transitions from rigid rotation to springy, buckling, and eventually complex deformation modes, unable to resist hydrodynamic stresses (Kanchan and Maniyeri, 2019) [9].
Figure 4. Deformation of a filament at different time points for various bending rigidity values (Kb), illustrating distinct deformation modes: rigid (Kb = 1.0), springy (Kb = 0.7), C-shape (Kb = 2.5 × 10−2), S-shape (Kb = 6 × 10−3), and complex deformations (Kb = 5 × 10−4). As bending rigidity decreases, the filament transitions from rigid rotation to springy, buckling, and eventually complex deformation modes, unable to resist hydrodynamic stresses (Kanchan and Maniyeri, 2019) [9].
Fluids 10 00217 g004
Figure 5. Classification of filament orbit regimes based on viscous flow forcing (VFF) and bending rigidity (Kb) for a fixed filament length L = 0.1 and shear rates K = 10, 16, 32. Results are compared with experimental works of Forgacs and Mason [10] and numerical studies of Stokie et al. [11]. Regimes are demarcated by solid vertical lines into rigid (175° < β < 180°), springy (90° < β < 175°), C-shape (30° < β < 90°), and complex (β < 30°) deformation classes using the exterior angle (β) (Kanchan and Maniyeri, 2019) [9].
Figure 5. Classification of filament orbit regimes based on viscous flow forcing (VFF) and bending rigidity (Kb) for a fixed filament length L = 0.1 and shear rates K = 10, 16, 32. Results are compared with experimental works of Forgacs and Mason [10] and numerical studies of Stokie et al. [11]. Regimes are demarcated by solid vertical lines into rigid (175° < β < 180°), springy (90° < β < 175°), C-shape (30° < β < 90°), and complex (β < 30°) deformation classes using the exterior angle (β) (Kanchan and Maniyeri, 2019) [9].
Fluids 10 00217 g005
Figure 6. Visualization of three geometries: (a) unit sphere, (b) ellipsoid with aspect ratio 1:2:4, (c) complex surface based on the real part of the spherical harmonic Y 3 2 (Lai and Seol, 2022) [38].
Figure 6. Visualization of three geometries: (a) unit sphere, (b) ellipsoid with aspect ratio 1:2:4, (c) complex surface based on the real part of the spherical harmonic Y 3 2 (Lai and Seol, 2022) [38].
Fluids 10 00217 g006
Figure 7. Snapshots of prolate vesicle motion under shear flow at different Reynolds numbers: (ae) Re = 1, (fj) Re = 5, (ko) Re = 9. The other parameters are ν = 0.9, Ca = 2.5, and λ = 20. These snapshots highlight transitions from tumbling to tank-treading motion, confirming that vesicle dynamics at moderate Re (Re = O(1)) are significantly influenced by fluid inertia (Lai and Seol, 2022) [38].
Figure 7. Snapshots of prolate vesicle motion under shear flow at different Reynolds numbers: (ae) Re = 1, (fj) Re = 5, (ko) Re = 9. The other parameters are ν = 0.9, Ca = 2.5, and λ = 20. These snapshots highlight transitions from tumbling to tank-treading motion, confirming that vesicle dynamics at moderate Re (Re = O(1)) are significantly influenced by fluid inertia (Lai and Seol, 2022) [38].
Fluids 10 00217 g007
Figure 8. Snapshots of vesicle motion in Poiseuille flow for different reduced volumes (ν) and channel aspect ratios (d/e). The vesicle deforms into a bullet shape for (ad) with ν = 0.99 and d/e = 1, a parachute shape for (eh) with ν = 0.91 and d/e = 1, and a croissant shape for (il) with ν = 0.973 and d/e = 1.2 (Lai and Seol, 2022) [38].
Figure 8. Snapshots of vesicle motion in Poiseuille flow for different reduced volumes (ν) and channel aspect ratios (d/e). The vesicle deforms into a bullet shape for (ad) with ν = 0.99 and d/e = 1, a parachute shape for (eh) with ν = 0.91 and d/e = 1, and a croissant shape for (il) with ν = 0.973 and d/e = 1.2 (Lai and Seol, 2022) [38].
Fluids 10 00217 g008
Figure 9. Industrial-intensive in-line mixer type DLM/S, designed and manufactured by INDAG Maschinenbau GmbH (Borsfleth, Germany), consists of multiple stages of star-shaped mixing elements functioning as rotor and stator. The stator elements are fixed to the housing, while the rotor elements are mounted on a shaft connected to an electrical motor. (Left): Top view of the mixer showing the N6 injector at the bottom. (Right): Magnified frontal view highlighting the rotor and stator arrangement, which ensures efficient mixing of highly viscous liquids with viscosities up to 70 kPa·s (Zargaran et al., 2023) [130].
Figure 9. Industrial-intensive in-line mixer type DLM/S, designed and manufactured by INDAG Maschinenbau GmbH (Borsfleth, Germany), consists of multiple stages of star-shaped mixing elements functioning as rotor and stator. The stator elements are fixed to the housing, while the rotor elements are mounted on a shaft connected to an electrical motor. (Left): Top view of the mixer showing the N6 injector at the bottom. (Right): Magnified frontal view highlighting the rotor and stator arrangement, which ensures efficient mixing of highly viscous liquids with viscosities up to 70 kPa·s (Zargaran et al., 2023) [130].
Fluids 10 00217 g009
Figure 10. Isometric view of the computational grid used for the mixer simulation, featuring over 2 million fully structured hexahedral cells. The grid is refined near critical zones of high shear gradients, such as rotor–stator gaps and the additive injection location, as shown in the magnified sections. An internal face highlighted in green marks the injection location for base phase tracer particles to accelerate simulation time, particularly for highly viscous mixtures. Stator mixing elements are excluded for improved clarity (Zargaran et al., 2023) [130].
Figure 10. Isometric view of the computational grid used for the mixer simulation, featuring over 2 million fully structured hexahedral cells. The grid is refined near critical zones of high shear gradients, such as rotor–stator gaps and the additive injection location, as shown in the magnified sections. An internal face highlighted in green marks the injection location for base phase tracer particles to accelerate simulation time, particularly for highly viscous mixtures. Stator mixing elements are excluded for improved clarity (Zargaran et al., 2023) [130].
Fluids 10 00217 g010
Figure 11. Velocity vectors for four different mixing scenarios at rotor angles of 0° and 30°, maintaining constant total mass flow rate. The emergence of eddies intensifies as the rotational Reynolds number ( R e r o t ) increases, visible between rotor stages and upstream of the rotor. Low R e r o t flows suppress eddies, limiting back-mixing. Simulations align with experimental observations in aqueous media, demonstrating longer micro-mixing times for higher viscosity flows (Zargaran et al., 2023) [130].
Figure 11. Velocity vectors for four different mixing scenarios at rotor angles of 0° and 30°, maintaining constant total mass flow rate. The emergence of eddies intensifies as the rotational Reynolds number ( R e r o t ) increases, visible between rotor stages and upstream of the rotor. Low R e r o t flows suppress eddies, limiting back-mixing. Simulations align with experimental observations in aqueous media, demonstrating longer micro-mixing times for higher viscosity flows (Zargaran et al., 2023) [130].
Fluids 10 00217 g011
Figure 12. Process for generating the nodal immersed boundary (NIB). The immersed geometry (Γ) is discretized into a triangulation. Gauss points from the elements of this triangulation are used to create particles. Subsequently, the triangulation is discarded and the simulation is performed using particles associated with the Gauss points of the immersed triangulation (Joachim et al., 2023) [132].
Figure 12. Process for generating the nodal immersed boundary (NIB). The immersed geometry (Γ) is discretized into a triangulation. Gauss points from the elements of this triangulation are used to create particles. Subsequently, the triangulation is discarded and the simulation is performed using particles associated with the Gauss points of the immersed triangulation (Joachim et al., 2023) [132].
Fluids 10 00217 g012
Figure 13. Slice of the fluid mesh at the beginning and the end of the simulation, demonstrating the use of adaptive meshing. (a) Initial coarse mesh at t = 0 s with uniform refinement ( p f l u i d = 4). (b) Refined mesh at t = 0.2 s using adaptive refinement ( p f l u i d = 7) for Re = 1200 (ν = 2 × 10 4 m2/s). The adaptive mesh approach reduces computational costs by allowing higher refinement near the impeller while maintaining fewer cells overall (Joachim et al., 2023) [132].
Figure 13. Slice of the fluid mesh at the beginning and the end of the simulation, demonstrating the use of adaptive meshing. (a) Initial coarse mesh at t = 0 s with uniform refinement ( p f l u i d = 4). (b) Refined mesh at t = 0.2 s using adaptive refinement ( p f l u i d = 7) for Re = 1200 (ν = 2 × 10 4 m2/s). The adaptive mesh approach reduces computational costs by allowing higher refinement near the impeller while maintaining fewer cells overall (Joachim et al., 2023) [132].
Fluids 10 00217 g013
Figure 14. Fluid velocity magnitude profile in the plane passing through the middle of the blades. The left image (a) represents flow at ν = 1.2 × 10−3 m2/s (Re ≈ 200) for a single impeller mixing rig, where the flow oscillates in the wake of the overlapping region between two impellers, with reduced velocity zones highlighted by black dashed ellipses. The right image (b) corresponds to ν = 2 × 10−4 m2/s (Re ≈ 1000), showing a more turbulent flow with multiple eddies propagating through the vessel, highlighted by black dashed circles. The velocity magnitude ranges from 0 (dark blue) to 7.4 m/s (tip-speed ratio of 1.25) (Joachim et al., 2023) [132].
Figure 14. Fluid velocity magnitude profile in the plane passing through the middle of the blades. The left image (a) represents flow at ν = 1.2 × 10−3 m2/s (Re ≈ 200) for a single impeller mixing rig, where the flow oscillates in the wake of the overlapping region between two impellers, with reduced velocity zones highlighted by black dashed ellipses. The right image (b) corresponds to ν = 2 × 10−4 m2/s (Re ≈ 1000), showing a more turbulent flow with multiple eddies propagating through the vessel, highlighted by black dashed circles. The velocity magnitude ranges from 0 (dark blue) to 7.4 m/s (tip-speed ratio of 1.25) (Joachim et al., 2023) [132].
Fluids 10 00217 g014
Figure 15. Aspects of (a) partitioned and (b) immersed FSI formulations are integrated in the (c) penalty–ILE coupling. The fluid–structure interface representation (determined by χ) conforms to the boundary of the structure (determined by ξ) in an approximate sense. The discrepancy between the two representations is exaggerated here for illustration purposes. In practice, we ensure that this maximum displacement is no greater than 0.1 of the Cartesian grid spacing. The local fluid velocity determines the motion of the surface representation Γ t f s , whereas the no-slip condition is satisfied in an approximate sense by spring-like forces that penalize displacements between the two representations of the fluid–structure interface (Kolahdouz et al., 2023) [129].
Figure 15. Aspects of (a) partitioned and (b) immersed FSI formulations are integrated in the (c) penalty–ILE coupling. The fluid–structure interface representation (determined by χ) conforms to the boundary of the structure (determined by ξ) in an approximate sense. The discrepancy between the two representations is exaggerated here for illustration purposes. In practice, we ensure that this maximum displacement is no greater than 0.1 of the Cartesian grid spacing. The local fluid velocity determines the motion of the surface representation Γ t f s , whereas the no-slip condition is satisfied in an approximate sense by spring-like forces that penalize displacements between the two representations of the fluid–structure interface (Kolahdouz et al., 2023) [129].
Fluids 10 00217 g015
Figure 16. Illustration of the penalty–ILE coupling for compressible structures. (a) Partitioned formulation and (b) immersed FSI formulations combine to form (c) the penalty–ILE coupling. The fluid–structure interface (χ) approximates the inner structural boundary (ξ), with exaggerated displacement shown for clarity (Kolahdouz et al., 2023) [129].
Figure 16. Illustration of the penalty–ILE coupling for compressible structures. (a) Partitioned formulation and (b) immersed FSI formulations combine to form (c) the penalty–ILE coupling. The fluid–structure interface (χ) approximates the inner structural boundary (ξ), with exaggerated displacement shown for clarity (Kolahdouz et al., 2023) [129].
Fluids 10 00217 g016
Figure 17. Snapshots of the computed velocity field and soft disk motion in a lid-driven cavity flow at (a) t = 0.0s, (b) t = 2.0 s, (c) t = 3.5 s, (d) t = 5.25 s, (e) t = 6.5 s, and (f) t = 9.0 s. The disk deformation and induced flow illustrate the strong interaction between the fluid and structure. Colors represent the velocity field, with near-contact dynamics emphasized at later times (Kolahdouz et al., 2023) [129].
Figure 17. Snapshots of the computed velocity field and soft disk motion in a lid-driven cavity flow at (a) t = 0.0s, (b) t = 2.0 s, (c) t = 3.5 s, (d) t = 5.25 s, (e) t = 6.5 s, and (f) t = 9.0 s. The disk deformation and induced flow illustrate the strong interaction between the fluid and structure. Colors represent the velocity field, with near-contact dynamics emphasized at later times (Kolahdouz et al., 2023) [129].
Fluids 10 00217 g017
Figure 18. Schematic representation of the deformed plate showing angular oscillations used for post-processing. The figure highlights the location of the probe point and the method for computing transverse displacement relative to the plate’s mean position to analyze modal contributions (Pandey et al., 2023) [123].
Figure 18. Schematic representation of the deformed plate showing angular oscillations used for post-processing. The figure highlights the location of the probe point and the method for computing transverse displacement relative to the plate’s mean position to analyze modal contributions (Pandey et al., 2023) [123].
Fluids 10 00217 g018
Figure 19. Instantaneous vorticity field and plate deformation at various time points for a plate with E = 1400 and M = 2, illustrating the flow-induced vibration response (Pandey et al., 2023) [123].
Figure 19. Instantaneous vorticity field and plate deformation at various time points for a plate with E = 1400 and M = 2, illustrating the flow-induced vibration response (Pandey et al., 2023) [123].
Fluids 10 00217 g019
Figure 20. (a) Variation in plate vibration amplitude A θ and dimensionless energy exchange (WWW) between fluid and solid with reduced velocity ( U R ) . Key regions illustrate transitions due to decreasing elasticity. (b) Contour plot of plate frequency ( f p ) as a function of reduced velocity ( U R ) .  The plot highlights the first ( f n 1 )- and second ( f n 2 )-mode natural frequencies, along with peak oscillation frequency ( f p P e a k ), lift frequency ( f L ), and drag frequency ( f D ). (c) Contributions of initial five Euler–Bernoulli natural modes (modes 1–5) to total energy ( E ¯ i ) with varying reduced velocity ( U R ) (Pandey et al., 2023) [123].
Figure 20. (a) Variation in plate vibration amplitude A θ and dimensionless energy exchange (WWW) between fluid and solid with reduced velocity ( U R ) . Key regions illustrate transitions due to decreasing elasticity. (b) Contour plot of plate frequency ( f p ) as a function of reduced velocity ( U R ) .  The plot highlights the first ( f n 1 )- and second ( f n 2 )-mode natural frequencies, along with peak oscillation frequency ( f p P e a k ), lift frequency ( f L ), and drag frequency ( f D ). (c) Contributions of initial five Euler–Bernoulli natural modes (modes 1–5) to total energy ( E ¯ i ) with varying reduced velocity ( U R ) (Pandey et al., 2023) [123].
Fluids 10 00217 g020
Figure 21. Computational domain consisting of a vertically held flexible plate in laminar viscous flow. The domain has dimensions of 20 H × 12 H. The plate, located at (4.6, 6.0), has a length H = 1.0 and thickness B = 0.01. Dirichlet boundary condition (U = 1.0) is applied at the inlet, and Neumann boundary condition is applied at the outlet. Top and bottom walls are subjected to no-slip conditions (Kanchan et al., 2024) [124].
Figure 21. Computational domain consisting of a vertically held flexible plate in laminar viscous flow. The domain has dimensions of 20 H × 12 H. The plate, located at (4.6, 6.0), has a length H = 1.0 and thickness B = 0.01. Dirichlet boundary condition (U = 1.0) is applied at the inlet, and Neumann boundary condition is applied at the outlet. Top and bottom walls are subjected to no-slip conditions (Kanchan et al., 2024) [124].
Fluids 10 00217 g021
Figure 22. Computational grids (shown in green) surrounding the flexible plate (highlighted in red) on the right side of the image (Kanchan et al., 2024) [124].
Figure 22. Computational grids (shown in green) surrounding the flexible plate (highlighted in red) on the right side of the image (Kanchan et al., 2024) [124].
Fluids 10 00217 g022
Figure 23. Flow-induced deformation response of the flexible plate, illustrating various branches from the study: low amplitude (LA), initial branch (IB), intermediate (I), upper branch (UB), bi-stable (BS), and lower branch (LB). (a) Normalized amplitude (A) versus reduced velocity U r ; (b) power spectral density (PSD) contours of vertical displacement (y) with normalized frequency and reduced velocity; (c) PSD contours for lift force coefficient ( C L ) with normalized frequency and reduced velocity (Kanchan et al., 2024) [124].
Figure 23. Flow-induced deformation response of the flexible plate, illustrating various branches from the study: low amplitude (LA), initial branch (IB), intermediate (I), upper branch (UB), bi-stable (BS), and lower branch (LB). (a) Normalized amplitude (A) versus reduced velocity U r ; (b) power spectral density (PSD) contours of vertical displacement (y) with normalized frequency and reduced velocity; (c) PSD contours for lift force coefficient ( C L ) with normalized frequency and reduced velocity (Kanchan et al., 2024) [124].
Fluids 10 00217 g023
Figure 24. 2D color contours of instantaneous spanwise vorticity in the wall-normal plane for clamped flexible filaments. Maximum positive (left) and negative (right) deflections of the filament centerline are shown (Agarwal et al., 2024) [134].
Figure 24. 2D color contours of instantaneous spanwise vorticity in the wall-normal plane for clamped flexible filaments. Maximum positive (left) and negative (right) deflections of the filament centerline are shown (Agarwal et al., 2024) [134].
Fluids 10 00217 g024
Figure 25. Comparison of the time evolution with experimental and numerical studies (Pinelli et al. [137]) of the streamwise tip displacement of the last filament in an oscillatory channel flow (Agarwal et al., 2024) [134].
Figure 25. Comparison of the time evolution with experimental and numerical studies (Pinelli et al. [137]) of the streamwise tip displacement of the last filament in an oscillatory channel flow (Agarwal et al., 2024) [134].
Fluids 10 00217 g025
Figure 26. Snapshots of the streamwise velocity contour and wavelike configurations of the filaments.
Figure 26. Snapshots of the streamwise velocity contour and wavelike configurations of the filaments.
Fluids 10 00217 g026
Figure 27. (Left) Illustration of the setup used in penalization for isothermal or T-neutral conditions. Fluids 10 00217 i001 Eulerian points (cell centers), Fluids 10 00217 i002 penalized cells. (Right) Illustration of the setup used in penalization for an adiabatic condition. Fluids 10 00217 i003 Solid interior cells, Fluids 10 00217 i004 ghost cells, • ghost point, Fluids 10 00217 i005 interface point, Fluids 10 00217 i006 first image point, Fluids 10 00217 i007 second image point, Fluids 10 00217 i008 Lagrangian points (Xu and Choi, 2023) [146].
Figure 27. (Left) Illustration of the setup used in penalization for isothermal or T-neutral conditions. Fluids 10 00217 i001 Eulerian points (cell centers), Fluids 10 00217 i002 penalized cells. (Right) Illustration of the setup used in penalization for an adiabatic condition. Fluids 10 00217 i003 Solid interior cells, Fluids 10 00217 i004 ghost cells, • ghost point, Fluids 10 00217 i005 interface point, Fluids 10 00217 i006 first image point, Fluids 10 00217 i007 second image point, Fluids 10 00217 i008 Lagrangian points (Xu and Choi, 2023) [146].
Fluids 10 00217 g027
Figure 28. Supersonic flow past a sphere (M = 2, Re = 300) (Xu and Choi, 2023) [146].
Figure 28. Supersonic flow past a sphere (M = 2, Re = 300) (Xu and Choi, 2023) [146].
Fluids 10 00217 g028
Figure 29. Temperature and vorticity contour at different Rayleigh numbers. Top (ad): temperature with a 0.05 increment; bottom (eh): vorticity; left to right: Ra = 103, 104, 105, 106 (Wang et al., 2023) [150].
Figure 29. Temperature and vorticity contour at different Rayleigh numbers. Top (ad): temperature with a 0.05 increment; bottom (eh): vorticity; left to right: Ra = 103, 104, 105, 106 (Wang et al., 2023) [150].
Fluids 10 00217 g029
Figure 30. The settling velocities of a single particle sedimentation: (a) without heat transfer; (b) with heat transfer at different Grashof numbers (Wang et al., 2023) [150].
Figure 30. The settling velocities of a single particle sedimentation: (a) without heat transfer; (b) with heat transfer at different Grashof numbers (Wang et al., 2023) [150].
Fluids 10 00217 g030
Figure 31. Particle locations at different times: (a) t = 10; (b) t = 20; (c) t = 30, where black, red, and blue indicate results for cases with Gr = 0, 100, and −100 (Wang et al., 2023) [150].
Figure 31. Particle locations at different times: (a) t = 10; (b) t = 20; (c) t = 30, where black, red, and blue indicate results for cases with Gr = 0, 100, and −100 (Wang et al., 2023) [150].
Fluids 10 00217 g031
Figure 32. Illustration of the immersed boundary method for Robin boundary conditions. The two yellow circles denote the auxiliary layers, and the blue line represents the solid boundary. The Robin boundary condition is imposed on the solid boundary, and the Lagrangian points on these three layers are distributed along the external normal direction (Wu et al., 2024) [152].
Figure 32. Illustration of the immersed boundary method for Robin boundary conditions. The two yellow circles denote the auxiliary layers, and the blue line represents the solid boundary. The Robin boundary condition is imposed on the solid boundary, and the Lagrangian points on these three layers are distributed along the external normal direction (Wu et al., 2024) [152].
Fluids 10 00217 g032
Figure 33. Comparisons between instantaneous vorticity field around a heaving airfoil provided by Martín-Alcántara et al. (left column) and the present results (middle column) at Re = 500, Sta = 0.16, and Stc = 0.5. The instantaneous isotherms around the heaving heated airfoil are presented in the right column. (ac) t/Tc = 1/8, (df) t/Tc = 1/4, (gi) t/Tc = 5/8, and (jl) t/Tc = 3/4. The instantaneous vorticity contours obtained by the present method agree well with those of Martín-Alcántara et al., which qualitatively verify that the proposed method can accurately capture the flow structures around complex moving boundaries (Wu et al., 2024) [152].
Figure 33. Comparisons between instantaneous vorticity field around a heaving airfoil provided by Martín-Alcántara et al. (left column) and the present results (middle column) at Re = 500, Sta = 0.16, and Stc = 0.5. The instantaneous isotherms around the heaving heated airfoil are presented in the right column. (ac) t/Tc = 1/8, (df) t/Tc = 1/4, (gi) t/Tc = 5/8, and (jl) t/Tc = 3/4. The instantaneous vorticity contours obtained by the present method agree well with those of Martín-Alcántara et al., which qualitatively verify that the proposed method can accurately capture the flow structures around complex moving boundaries (Wu et al., 2024) [152].
Fluids 10 00217 g033
Figure 34. Numerical accuracy test of the proposed EIB-RTLBFS by the transient heat diffusion of a Gaussian hill. (a) Comparison of the isotherms of predicted results (red dashed line) and analytical solution (blue solid line) at Fo = 0.005. (b) Convergence of numerical error versus mesh spacing for unsteady heat diffusion of a Gaussian hill, indicating that the overall accuracy of the EIB-RTLBFS is of second order. The solidlines with symbols indicate results obtained at each interval and dashed line indicates a line having slope of 2. (Ménez et al., 2023) [172].
Figure 34. Numerical accuracy test of the proposed EIB-RTLBFS by the transient heat diffusion of a Gaussian hill. (a) Comparison of the isotherms of predicted results (red dashed line) and analytical solution (blue solid line) at Fo = 0.005. (b) Convergence of numerical error versus mesh spacing for unsteady heat diffusion of a Gaussian hill, indicating that the overall accuracy of the EIB-RTLBFS is of second order. The solidlines with symbols indicate results obtained at each interval and dashed line indicates a line having slope of 2. (Ménez et al., 2023) [172].
Fluids 10 00217 g034
Figure 35. Instantaneous (a) vorticity structures (visualized by the iso-surface of Q = 0.2), (b) iso-surface of T = 0.1 colored by the streamwise velocity, and (c) isotherm on the z-mid plane for the heated flapping membrane structure at a singular moment and Re = 200; (d) side view of the flapping profile of the heated membrane structure (Ménez et al., 2023) [172].
Figure 35. Instantaneous (a) vorticity structures (visualized by the iso-surface of Q = 0.2), (b) iso-surface of T = 0.1 colored by the streamwise velocity, and (c) isotherm on the z-mid plane for the heated flapping membrane structure at a singular moment and Re = 200; (d) side view of the flapping profile of the heated membrane structure (Ménez et al., 2023) [172].
Fluids 10 00217 g035
Figure 36. Schematic plot of a droplet’s impact on a circular cylinder (Niu et al., 2022) [191].
Figure 36. Schematic plot of a droplet’s impact on a circular cylinder (Niu et al., 2022) [191].
Fluids 10 00217 g036
Figure 37. Shape evolution and velocity field for a droplet impacting on a hydrophilic cylinder surface with θ = 60°, Re = 700, We = 10 (Niu et al., 2022) [191].
Figure 37. Shape evolution and velocity field for a droplet impacting on a hydrophilic cylinder surface with θ = 60°, Re = 700, We = 10 (Niu et al., 2022) [191].
Fluids 10 00217 g037
Figure 38. Computational domain for plane waves incident on two spherical bubbles (side view) (Hou et al., 2023) [196].
Figure 38. Computational domain for plane waves incident on two spherical bubbles (side view) (Hou et al., 2023) [196].
Fluids 10 00217 g038
Figure 39. Interior velocity and pressure. Larger bubbles of different radii and a smaller bubble of fixed size (fi = 2437 Hz, P0  = 106 Pa, d = 3a = 0.03 m. (a) z-axis. (b) Bubbles in contact (case 7, compression). (c) Case 7, expansion. (d) Case 8, spherical and prolate (expansion) (Hou et al., 2023) [196].
Figure 39. Interior velocity and pressure. Larger bubbles of different radii and a smaller bubble of fixed size (fi = 2437 Hz, P0  = 106 Pa, d = 3a = 0.03 m. (a) z-axis. (b) Bubbles in contact (case 7, compression). (c) Case 7, expansion. (d) Case 8, spherical and prolate (expansion) (Hou et al., 2023) [196].
Fluids 10 00217 g039
Figure 40. (a) Computational domain, (b) grid resolution, (c) and extent of yielded (blue) and unyielded (red) regions around solid objects (Fazli et al., 2023) [207].
Figure 40. (a) Computational domain, (b) grid resolution, (c) and extent of yielded (blue) and unyielded (red) regions around solid objects (Fazli et al., 2023) [207].
Fluids 10 00217 g040
Figure 41. Pressure and viscosity fields for flow over a fixed sphere at Re′ = 0.002956 and Bi* = 0.747 (Fazli et al., 2023) [207].
Figure 41. Pressure and viscosity fields for flow over a fixed sphere at Re′ = 0.002956 and Bi* = 0.747 (Fazli et al., 2023) [207].
Fluids 10 00217 g041
Figure 42. Pressure and viscosity fields for flow over a fixed sphere at Re′ = 0.97 and Bi* = 0.108 (Fazli et al., 2023) [207].
Figure 42. Pressure and viscosity fields for flow over a fixed sphere at Re′ = 0.97 and Bi* = 0.108 (Fazli et al., 2023) [207].
Fluids 10 00217 g042
Figure 43. A priori known geometry incorporated by the indicator function α with an unknown void reconstructed by the scaling function γ. Blue: physical domain, red: unknown void domain, white: fictitious domain (Bürchner et al., 2023) [206].
Figure 43. A priori known geometry incorporated by the indicator function α with an unknown void reconstructed by the scaling function γ. Blue: physical domain, red: unknown void domain, white: fictitious domain (Bürchner et al., 2023) [206].
Fluids 10 00217 g043
Figure 44. 1D interface problem p = 2. The dashed black vertical lines visualize the boundaries of the cut element and the red vertical line the material interface (Bürchner et al., 2023) [206].
Figure 44. 1D interface problem p = 2. The dashed black vertical lines visualize the boundaries of the cut element and the red vertical line the material interface (Bürchner et al., 2023) [206].
Fluids 10 00217 g044
Figure 45. ρ -scaling: reconstructed scaling function γ for an elliptical void. The dotted white ellipse shows the position and size of the real elliptical void (Bürchner et al., 2023) [206].
Figure 45. ρ -scaling: reconstructed scaling function γ for an elliptical void. The dotted white ellipse shows the position and size of the real elliptical void (Bürchner et al., 2023) [206].
Fluids 10 00217 g045
Figure 46. Two IBM particles near a wall over a Cartesian uniform collocated grid (Eulerian framework). Lagrangian markers are shown with filled circles. The non-modified and modified interpolation operator’s support at the position of the Lagrangian markers are shown in blue and black, respectively (Chéron et al., 2023) [208].
Figure 46. Two IBM particles near a wall over a Cartesian uniform collocated grid (Eulerian framework). Lagrangian markers are shown with filled circles. The non-modified and modified interpolation operator’s support at the position of the Lagrangian markers are shown in blue and black, respectively (Chéron et al., 2023) [208].
Fluids 10 00217 g046
Figure 47. Instantaneous velocity field of the random arrays of monodispersed spheres at solid volume fraction 50% at t = 1.5 s for the standard MLS–IBM and HyBM (left and right, respectively) for two numerical resolutions, Dp/Δx = 16 and Dp/Δx = 32 (top and bottom row, respectively) (Chéron et al., 2023) [208].
Figure 47. Instantaneous velocity field of the random arrays of monodispersed spheres at solid volume fraction 50% at t = 1.5 s for the standard MLS–IBM and HyBM (left and right, respectively) for two numerical resolutions, Dp/Δx = 16 and Dp/Δx = 32 (top and bottom row, respectively) (Chéron et al., 2023) [208].
Fluids 10 00217 g047
Figure 48. General experimental and numerical setup of the fixed packed bed reactor with 139 spherical particles in body-centered cubic packing (BCC). For the experiments, the packing inside the bulk reactor (R) is illuminated by a laser light sheet (L) and the measurement signal is recorded by the camera (C) (Gorges et al., 2024) [216].
Figure 48. General experimental and numerical setup of the fixed packed bed reactor with 139 spherical particles in body-centered cubic packing (BCC). For the experiments, the packing inside the bulk reactor (R) is illuminated by a laser light sheet (L) and the measurement signal is recorded by the camera (C) (Gorges et al., 2024) [216].
Fluids 10 00217 g048
Figure 49. Contour plots of the y-component of the mean velocity vector field in the z = 0.04309 m plane and a height of y = 0.495 m to y = 0.54 m for Rep = 300 for three different dpx ratios. The white box in the plots of the smooth and blocked-off IBM simulation results corresponds to the measurement area in the experimental results (Gorges et al., 2024) [216].
Figure 49. Contour plots of the y-component of the mean velocity vector field in the z = 0.04309 m plane and a height of y = 0.495 m to y = 0.54 m for Rep = 300 for three different dpx ratios. The white box in the plots of the smooth and blocked-off IBM simulation results corresponds to the measurement area in the experimental results (Gorges et al., 2024) [216].
Fluids 10 00217 g049
Figure 50. Qualitative cutout of the simulation mesh for the smooth IBM for the z-plane situated at z = 0.04309 m for illustration of the refined mesh for the front-position measuring plane. The bottom zoomed area shows the refined mesh at the velocity-inlet at the bottom of the reactor and the second zoomed area shows the refined mesh around the last layers of particles and the beginning of the region of interest (Gorges et al., 2024) [216].
Figure 50. Qualitative cutout of the simulation mesh for the smooth IBM for the z-plane situated at z = 0.04309 m for illustration of the refined mesh for the front-position measuring plane. The bottom zoomed area shows the refined mesh at the velocity-inlet at the bottom of the reactor and the second zoomed area shows the refined mesh around the last layers of particles and the beginning of the region of interest (Gorges et al., 2024) [216].
Fluids 10 00217 g050
Figure 51. Particle trajectories for two falling particles. The gravity vector acts in the positive x-direction, and thus particles are moving from left to right. P1 is the heavier particle (Giahi and Bergstrom, 2023) [222]. The numerical data is adopted from Uhlmann [225].
Figure 51. Particle trajectories for two falling particles. The gravity vector acts in the positive x-direction, and thus particles are moving from left to right. P1 is the heavier particle (Giahi and Bergstrom, 2023) [222]. The numerical data is adopted from Uhlmann [225].
Fluids 10 00217 g051
Figure 52. Position and velocity of two falling particles: (a) particle x location, (b) particle x--component of velocity, (c) particle y location, and (d) particle y-component of velocity (Giahi and Bergstrom, 2023) [222]. The numerical data is adopted from Uhlmann [225].
Figure 52. Position and velocity of two falling particles: (a) particle x location, (b) particle x--component of velocity, (c) particle y location, and (d) particle y-component of velocity (Giahi and Bergstrom, 2023) [222]. The numerical data is adopted from Uhlmann [225].
Fluids 10 00217 g052
Figure 53. (a) Angular position and (b) angular velocity of the two falling particles (Giahi and Bergstrom, 2023) [222]. The numerical data is adopted from Uhlmann [225].
Figure 53. (a) Angular position and (b) angular velocity of the two falling particles (Giahi and Bergstrom, 2023) [222]. The numerical data is adopted from Uhlmann [225].
Fluids 10 00217 g053
Figure 54. Normalized force on the plate during the first six oscillation cycles of oscillating in oscillating and orbital flows. (a) KC = 0.5. (b) KC = 1.5. (c) KC = 2.5. Present, oscillating: present results for oscillating flow. Present, orbital: present results for orbital flow (Yu et al., 2023) [127]. CFD results for orbital flow simulated by Mentzoni and Kristiansen [226].
Figure 54. Normalized force on the plate during the first six oscillation cycles of oscillating in oscillating and orbital flows. (a) KC = 0.5. (b) KC = 1.5. (c) KC = 2.5. Present, oscillating: present results for oscillating flow. Present, orbital: present results for orbital flow (Yu et al., 2023) [127]. CFD results for orbital flow simulated by Mentzoni and Kristiansen [226].
Fluids 10 00217 g054
Figure 55. Zoomed-in plot of the streamlines close to the solid part of the perforated plate underneath the free surface. The mean water surface is at Z = 0. The plot is a snapshot at t/T = 11.54 (t = 15 s). Other considered parameters are KC = 0.50 and kD = 0.87 (T = 1.3 s). (a) Z/D = 1.67 (Z = −0.6 m); (b) Z/D = 0.83 (Z = −0.3 m); (c) Z/D = 0.28 (Z = −0.1 m) (Yu et al., 2023) [127].
Figure 55. Zoomed-in plot of the streamlines close to the solid part of the perforated plate underneath the free surface. The mean water surface is at Z = 0. The plot is a snapshot at t/T = 11.54 (t = 15 s). Other considered parameters are KC = 0.50 and kD = 0.87 (T = 1.3 s). (a) Z/D = 1.67 (Z = −0.6 m); (b) Z/D = 0.83 (Z = −0.3 m); (c) Z/D = 0.28 (Z = −0.1 m) (Yu et al., 2023) [127].
Fluids 10 00217 g055
Figure 56. Zoomed-in plot of the vorticity field close to the solid part of the perforated plate underneath the free surface. (a) KC = 0.29. (b) KC = 0.50. The mean water surface is at Z = 0. The plot is a snapshot at t/T = 11.54 (15 s). Other considered parameters are non-dimensional wave number kD = 0.87 (T = 1.3 s) and non-dimensional submergence Z/D = 0.28 (Z = −0.1 m) (Yu et al., 2023) [127].
Figure 56. Zoomed-in plot of the vorticity field close to the solid part of the perforated plate underneath the free surface. (a) KC = 0.29. (b) KC = 0.50. The mean water surface is at Z = 0. The plot is a snapshot at t/T = 11.54 (15 s). Other considered parameters are non-dimensional wave number kD = 0.87 (T = 1.3 s) and non-dimensional submergence Z/D = 0.28 (Z = −0.1 m) (Yu et al., 2023) [127].
Fluids 10 00217 g056
Table 1. Comparison between diffuse and sharp-interface immersed boundary methods.
Table 1. Comparison between diffuse and sharp-interface immersed boundary methods.
Feature/AspectDiffuse-Interface MethodSharp-Interface Method
Interface RepresentationInterface is smeared over a finite thickness (regularized)Interface is exact, represented as a discontinuity (zero-thickness)
Governing Equation ModificationEquations are modified over a band using indicator (phase) functionsGoverning equations remain valid, but interface conditions are sharply enforced
Interface TrackingUses phase field, level set, or volume-of-fluid (VOF) methodsUses body-fitted mesh, ghost-cell, or cut-cell methods
Force ApplicationDistributed body force (e.g., penalty or smoothing force) over the interface regionDiscrete delta function or sharp condition imposition at interface
Continuity TreatmentAllows for continuous transition in material properties like density or viscosityMaterial properties jump discontinuously at the interface
Numerical ComplexityEasier to implement on fixed grids (Eulerian); smooth solutionsRequires special treatment for jumps, boundary reconstruction, and interpolation
Mass ConservationCan suffer from slight mass loss/gain (especially in level-set based methods)Better mass conservation due to explicit boundary enforcement
UtilityFlows with complex moving interfaces, topological changes (e.g., breakup, coalescence)Problems requiring high accuracy at interface (e.g., fluid–structure interaction)
Table 2. Recent advancements in IBM considering boundary treatment and corresponding fluid–structural solvers for biological applications.
Table 2. Recent advancements in IBM considering boundary treatment and corresponding fluid–structural solvers for biological applications.
ReferenceFlow–
Structural Solver
Boundary TreatmentSummaryApplication AreaYear
Griffith et al., 2007 [13]FDMDiffuseAdaptive, second-order accurate IBM for heart valve simulationsCardiovascular (heart valves)2007
Heys et al., 2008 [14]FDMDiffuseModels filiform hair motion with viscous couplingArthropod sensory systems2008
Kim and Lai, 2010 [15]FDMDiffuseSimulates inextensible vesicles in fluid flowVesicle dynamics2010
Maniyeri et al., 2012 [16]FVMDiffuseBacterial flagellum propulsion in viscous fluidMicroorganism locomotion2012
Maniyeri and Kang, 2014 [17]FVMDiffuseBacterial flagellar bundling and tumblingMicroorganism locomotion2014
De Rosis, 2014 [18] LBMDiffuseTandem flapping wings with phase difference effectsInsect flight dynamics2014
Battista et al., 2018 [19]FDMDiffuseImplementation of IBM as 2D softwareBiological flows2018
Fai and Rycroft, 2018 [20]FEMDiffuseImproved accuracy for thin fluid layers in FSIVesicle migration in narrow channels2018
Kanchan and Maniyeri, 2019 [9] FVMDiffuseBuckling and recuperation of flexible filaments in shear flowDiatom chain dynamics2019
Kanchan and Maniyeri, 2020 [21]FVMDiffuseAsymmetric filament deformation in oscillatory flowFlexible filament dynamics2020
Kanchan and Maniyeri, 2020 [22]FVMDiffuseSelf-excited oscillation of filaments in channel flowFluid–structure interaction2020
Kanchan and Maniyeri, 2020 [23]FVMDiffuseMultiple-filament interaction in shear flowFilament dynamics2020
Wang et al., 2020 [24]FDMDiffuseSelf-propelled flexible plate with Navier slipBio-inspired propulsion2020
Meng et al., 2020 [25]LBMDiffuseDendritic growth and motion in convectionSolidification dynamics2020
Delong et al., 2014 [26]FVMDiffuseFluctuating immersed boundary for Brownian dynamicsStatic and Dynamic particle interaction2014
Coclite et al., 2020 [27]LBMDiffuseDynamic IBM for rigid/deformable objects in 3D flowGeneral FSI2020
Ong et al., 2021 [28]FDMDiffuseInextensible vesicles in 2D Stokes flowVesicle dynamics2021
Ghosh, 2021 [29]FDMDiffuseFluid-induced deformation of cantilever beamsBiofilm-fluid interaction2021
Casquero et al., 2021 [30]FEMDiffuseCapsule/vesicle dynamics with B-splines/T-splinesBiological membranes2021
Lampropoulos et al., 2021 [31]FEMDiffuseEfficient hemodynamic simulation in aneurysmsIntracranial aneurysm flow2021
Mirfendereski and Park, 2021 [32]FDMDiffusePulsatile flow in stenotic channelsBlood flow in arteries2021
Eldoe et al., 2022 [33]FVMDiffuseRigid filament interaction in oscillatory flowLow-Re filament dynamics2022
Kassen et al., 2022 [34]FDMDiffuseCell–cell interactions in whole bloodBlood cell dynamics2022
Ntetsika and Papadopoulos, 2022 [35]FDMDiffuseFilament dynamics in shear flow with ANN predictionFlexible filament behavior2022
Maniyeri, 2022 [36]FVMDiffuseElastic rod dynamics in fluid flowFlagellar motion2022
Zhu et al., 2022 [37] LBMDiffuseTea shoot deformation under negative pressureAgricultural robotics2022
Lai and Seol, 2022 [38]FDMDiffuseStable vesicle dynamics in 3D Navier–StokesVesicle dynamics2022
Bourantas et al., 2023 [39]FEMDiffuseEfficient blood flow simulation in complex vasculatureVascular hemodynamics2023
Kaiser et al., 2023 [40]FDMDiffuseHeart valve hemodynamics compared with 4D flow MRICardiovascular flow2023
Ladiges et al., 2022 [41]MACDiffuseSimulation of electrolytes in presence of physical boundariesInduced charge electro-osmosis2022
Luo et al., 2008 [42]FDMSharpNovel IBM for phonation and FSI in biological systemsVocal fold dynamics2008
Bhardwaj and Mittal, 2012 [12]FDM–FEMSharpCoupled IBM–FEM solver for large-scale flow-induced deformationFlexible structures in flow2012
Bhardwaj et al., 2013 [43]FDM–FEMSharpBlast-induced eye deformation using FSIBiomechanics (eye trauma)2013
Bailoor et al., 2017 [44]FDM–FEMSharpCompressible FSI with large deformations (ghost-cell method)Blast–structure interaction2017
Bourantas et al., 2021 [45]FEMSharpInternal flows in complex geometries (blood flow)Cardiovascular hemodynamics2021
Brown et al., 2022 [46]FDMSharpTAVR device simulation with patient-specific anatomyTranscatheter aortic valve replacement2022
Wang et al., 2022 [47]FEMSharp (moving)Heart valve flow simulation with hybrid methodCardiovascular flow2022
Singh and Kumar, 2023 [48]FVMSharpTumor morphology modeling without body-conformal gridsBioheat transfer in tumors2023
Table 3. Recent advancements in IBM considering boundary treatment and corresponding fluid solvers for VIV and flexible body interactions.
Table 3. Recent advancements in IBM considering boundary treatment and corresponding fluid solvers for VIV and flexible body interactions.
ReferenceFluid SolverBoundary TreatmentSummaryApplication AreaYear
Kumar et al., 2015 [57]LBMDiffuseClap-and-fling mechanism at low ReInsect flight2015
De Rosis, 2015 [58]LBMDiffuseTandem flapping wings near ground effectAerial vehicle design2015
Wang et al., 2018 [59]FDMDiffuseOscillating hydrofoils for energy harvestingHydrodynamics2018
Li et al., 2018 [60]FDMDiffusePitching motion profiles for energy extractionSemi-active foils2018
Xie et al., 2019 [61]FVMDiffuseVIV of a cylinder with attached filamentVortex-induced vibrations2019
Ma et al., 2019 [62]RANSDiffuseFlow physics within blade passagesTurbomachinery2019
Wu et al., 2019 [63]LBMDiffuseFlow-induced vibration of tandem elliptical cylindersTandem cylinder flows2019
Wang et al., 2019 [64]FVMDiffuseElastic bodies in viscous fluids (vegetation/turbines)FSI in laminar/turbulent flows2019
Wang et al., 2020 [24]FDM–FEMDiffuseHydrodynamic interaction of flexible flagsPoiseuille flow2020
Chen et al., 2020 [65]FDMDiffuse3D flag flapping dynamics in Poiseuille flowFlexible structures2020
Zhao et al., 2020 [66]FVMDiffuseTsunami-like wave impact on coastal bridgesCoastal engineering2020
Zhang et al., 2020 [67]FDMDiffuseHydrodynamics of tuna caudal keelsBio-inspired propulsion2020
Luo et al., 2021 [68] FDMDiffuseReduced-order FSI for biofluid systemsBiofluid dynamics2021
Kasbaoui et al. 2021 [69]DNSDiffuseSwirling von Kármán flow with moving IBMTurbulence transitions2021
Dong et al., 2021 [70]SGKSDiffuseChordwise deformation effects on batoid fishBio-inspired propulsion2021
Ai et al., 2021 [71]FDMDiffuseInternal wave prediction with IBMStratified flows2021
Yan et al., 2021 [72]LBMDiffuseVortex shedding around circular cylindersIsolated/tandem cylinders2021
Tian et al., 2021 [73]FEMDiffuseTransient FSI with energy conservationGeneral FSI2021
Zhao et al., 2021 [74]FDMDiffuseDrag reduction using micro floating raftsUnderwater vehicles2021
Yaswanth and Maniyeri, 2022 [75]FVMDiffuseFluid mixing in oscillating lid-driven cavityMixing enhancement2022
Mazharmanesh et al., 2022 [76]LBMDiffuseEnergy harvesting in inverted piezoelectric flagsPiezoelectric energy2022
Karimnejad et al., 2022 [77]LBMDiffusePulsating flow effects on particle settlingParticle dynamics2022
Mao et al., 2022 [78]FDMDiffuseDrag reduction with flexible hairy coatingsDrag reduction2022
Zhang et al., 2022 [79]DNSDiffuseHydrodynamics of tuna finletsBio-inspired propulsion2022
Yu and Yu, 2022 [80]FDMDiffuseFlow interactions with aquatic vegetationEnvironmental flows2022
Jin et al., 2022 [81]FVMDiffuseVOF-IBM for solitary wave free surface flowFree-surface flows2022
Fang et al., 2022 [82]LBMDiffuseFish swimming hydrodynamicsBio-inspired robotics2022
Huang et al., 2022 [83]LBMDiffuseFSI of fish moving through turbinesFish-turbine interactions2022
Xiao et al., 2022 [84] LBMDiffuseWater entry/exit of rigid bodiesFree-surface FSI2022
Mi et al., 2022 [85]URANSDiffuseFSI in submerged netsAquaculture nets2022
Stival et al., 2022 [86]FVMDiffuseLES–IBM for wind turbine flowsWind energy2022
Mazharmanesh et al., 2023 [87]LBMDiffuseTandem/side-by-side piezoelectric flags in oscillating flow; identified chaotic/symmetric regimesEnergy harvesting (inverted flags)2023
Luo and Zhang, 2023 [88]HydroFlowDiffuseIBM in σ-coordinate model for wave-structure interactionsOcean engineering2023
Dong and Huang, 2023 [89]LESDiffuseWall-modeled LES–IBM for batoid fish swimming; revealed hairpin vorticesBio-inspired propulsion2023
Guo and Hou, 2023 [90]Other (DUGKS)DiffuseSlip-condition IBM for AUV drag reductionUnderwater vehicles2023
Wu and Guo, 2023 [91]LBMDiffuseFS-LBM with IB for aircraft ditching; included surface tension effectsMultiphase FSI (water impact)2023
Park et al., 2023 [92]FDMDiffuseSemi-Lagrangian Navier–Stokes with reduced IBM for high-inertia/elasticity FSIGeneral FSI2023
Monteiro and Mariano, 2023 [93]SpectralDiffuseFourier pseudo-spectral-IBM for airfoils/VAWTs; validated at Re = 1000Wind energy2023
Liu et al., 2023 [94]LBMDiffuseIB–LBM with AMR/VOF for free-surface flows in ocean engineeringOcean engineering2023
Zhang et al., 2024 [95]FDMDiffusePerforated plate-filament system; observed MFD mode and drag reductionDrag reduction2024
Mao et al., 2023 [96]FDMDiffuse (Penalty IBM)Snap-through dynamics of buckled filament; mode transitions with varying bending rigidity/ReFlexible filaments in flow2023
Borazjani and Sotiropoulos, 2009 [97]FEMSharpVIV of two elastically mounted cylinders at Re = 200Vortex-induced vibrations2009
Seo and Mittal, 2011 [98]FDM–FEMSharpCut-cell IBM to reduce spurious pressure oscillationsMoving boundary flows2011
Griffith et al., 2016 [99]FDM–FEMSharpFlow around rotationally oscillating cylindersEnergy harvesting2016
Griffith and Leontini, 2017 [100]FDM–FEMSharpHeuristic model for VIV simulationsVortex-induced vibrations2017
Khalili et al., 2018 [101]FDMSharpHigh-order ghost-point IBM for compressible flowsCompressible viscous flows2018
Mishra et al., 2019 [102]FDM–FEMSharpViscoelastic plate attached to a cylinderFlow-induced oscillations2019
Majumdar et al., 2020 [103]FVMSharpDynamic transitions in plunging foil flowTransition to chaos2020
Narváez et al., 2020 [104]Incompact3DSharpFlow-induced vibrations of tandem cylindersTandem cylinder FIV2020
Xu et al., 2020 [105]FDMSharpHigh-order solutions for nonlinear water wavesFree-surface flows2020
Xin et al., 2020 [106]CIRSharpGhost cell method for sloshing in tanksFree-surface flows2020
Kundu et al., 2020 [107]FDM–FEMSharpFSI solver coupling IBM–FEM with dynamic under-relaxationGeneral FSI2020
Kwon et al., 2020 [108]FVMSharpShallow water flow around cylindersTsunami mitigation2020
Tsai and Lo, 2020 [109]FDMSharpFluid–structure interaction in submerged netsAquaculture nets2020
Seshadri and De, 2021 [110]FVMSharpRobust framework for moving bodiesGeneral FSI2021
Robaux and Benoît, 2021 [111]FDMSharpFully nonlinear potential wave tankFree-surface flows2021
Badhurshah et al., 2021 [112]FDM–FEMSharpVIV of a cylinder with bistable springsEnergy harvesting2021
Tong et al., 2021 [113]FDMSharpNonlinear wave-structure interactionsOffshore engineering2021
Hanssen and Greco, 2021 [114]FDMSharpOverlapping grid method for water wavesWave-body interactions2021
Sharma et al., 2022 [115]FDM–FEMSharpFIV of cylinders with varying cross-sectionsFlow-induced vibrations2022
Giannenas et al., 2022 [116]Incompact3DSharpWake response to harmonic forcingBluff body flows2022
Khedkar and Bhalla, 2022 [117]FDMSharpMPC-IBM for wave energy convertersRenewable energy2022
Song et al., 2022 [118]RANSSharp3D scour model for complex structuresCoastal/scour dynamics2022
Gómez et al., 2022 [119]Incompact3DSharpVIV in four circular cylindersMulti-cylinder FSI2022
Badhurshah et al., 2022 [120]FDM–FEMSharpVIV with bistable springs (energy harvesting)Vortex-induced vibrations2022
Ji et al., 2022 [121]LESSharpHybrid actuator line-IB for wind turbine wakesWind energy2022
Xu et al., 2023 [122]FDMSharpHigh-order finite difference solver with IBM for wave loads on marine structuresOffshore structures2023
Pandey et al., 2023 [123]FDM–FEMSharpFIV of cantilevered plate; identified lock-in regimes for energy harvestingEnergy harvesting2023
Kanchan et al., 2024 [124]FDM–FEMSharpANN-predicted oscillations of elastic plate in flow; 6 response regionsLow-Re FSI2024
Chern et al., 2023 [125]LESSharpPlasma-actuated dynamic stall control on flapping NACA 0012 wingAerodynamics (flow control)2023
Yang et al., 2023 [126]LESSharp (ghost cell)Modified IBM for VAWT simulations; validated with Eppler387 airfoilWind energy2023
Yu et al., 2023 [127]FDMSharp (IB-GHPC)2D wave tank with pressure-velocity separation; validated for free-surface flowsCoastal engineering2023
Sundar et al., 2023 [128]ALESharpMoving-boundary-enabled standard Navier–Stokes based Physics Informed Neural NetworkFlow past plunging foils2023
Kolahdouz et al., 2023 [129]FDM–FEMSharp (ILE)Dirichlet–Neumann coupling for nonlinear FSI; improved volume conservationLarge-deformation FSI2023
Zargaran et al., 2023 [130]FVM-FEMSharpLagrangian-IBM for rotor-stator mixers; reduced numerical diffusionIndustrial mixing2023
Li et al., 2023 [131]FDMSharp (level-set)3D sharp-interface IBM for turbomachinery flowsHydraulic turbomachinery2023
Joachim et al., 2023 [132]LetheSharp (Nitsche’s)Parallel NIB method for fluid mixing in stirred tanksIndustrial mixing2023
Kou and Ferrer, 2023 [133]FDMSharp (volume penalization)Combined volume penalization–SFD for moving geometries; damped spurious wavesMoving-boundary flows2023
Agarwal et al., 2024 [134]FDMSharp (direct forcing)IGA–FDM coupling for slender rods in flow; decoupled fluid-grid resolutionFlexible structures2024
Table 4. Recent advancements in IBM considering boundary treatment and corresponding fluid solvers for heat transfer applications.
Table 4. Recent advancements in IBM considering boundary treatment and corresponding fluid solvers for heat transfer applications.
ReferenceFluid SolverBoundary TreatmentSummaryApplication AreaYear
Ren et al., 2013 [138]FDMDiffuseDirichlet (velocity) + Neumann (heat flux) corrections for momentum/energy equationsConjugate heat transfer2013
Mazharmanesh et al., 2020 [139]LBMDiffuseMultiple piezoelectric flags in tandem/side-by-side configurationsEnergy harvesting2020
Tong et al., 2020 [140]LBMDiffuseMultiblock LBM–IBM for fouling/ash deposition in heat exchangersFouling dynamics2020
Abaszadeh et al., 2022 [141]LBMDiffuseIB–LBM for radiative heat transfer in 2D irregular geometriesRadiative transfer2022
Jiang et al., 2022 [142]LBMDiffuseParallel IB–LBM for fully resolved particle-laden flowsSuspension dynamics2022
Tao et al., 2022 [143]LBMDiffuseSimplified IB–LBM for thermal flows with no-slip/temperature BCsThermal flows2022
Haeri and Shrimpton, 2013 [144]FEMDiffuse (fictitious)Implicit fictitious domain method for flow–heat transfer past immersed objectsFluid-immersed object interaction2013
Chen et al., 2020 [145]LBMDiffuse (IB–STLBM)Simplified LBM with IBM for incompressible thermal flowsThermal flows with immersed objects2020
Xu and Choi, 2023 [146]FDMDiffuse (MIBPM)Monolithic IB projection method for incompressible flows with heat transferIncompressible thermal flows2023
Hosseini et al., 2021 [147]LBMDiffuse (MRT)IB–LBM for elastic vortex generator (EVG) in microchannelsMicroscale mixing2021
Wu et al., 2023 [148]LBMDiffuseExplicit IBM for TFSI with Neumann BCs (heat flux)Thermal-fluid–structure interaction2023
Chen et al., 2020 [149]FDMDiffuse (penalty IBM)Fluid–structure–thermal interaction of flexible flags in heated channelsHeat transfer enhancement2020
Wang et al., 2023 [150]FDM–FEMDiffuse
(penalty)
IBM for rarefied gas flow with slip-modeled velocity/temperature jumpsRarefied gas flows2023
Zhang et al., 2016 [151]LBMDiffuse (PIBM)Particulate IBM for thermal particle–fluid interactions with DEMParticle-laden flows2016
Wu et al., 2024 [152]LBMDiffuseImplicit IBM for TFSI with Robin BCs (combined convection–radiation)Thermal-fluid–structure interaction2024
Pacheco et al., 2005 [153]FVMSharpFinite-volume non-staggered grid with Dirichlet/Neumann BCs for complex geometriesGeneral heat transfer2005
Soti et al., 2015 [154]FDM–FEMSharpStrongly coupled FSI solver for flexible structures with convective heat transferEnergy harvesting2015
Garg et al., 2018 [155]FDM–FEMSharpVIV of cylinder with thermal buoyancy (Boussinesq approximation)Vortex-induced vibrations2018
Garg et al., 2020 [156]FDM–FEMSharpThermal buoyancy effects on wake-induced vibration (WIV) of square prismsBluff body heat transfer2020
Lou et al., 2020 [157]FVMSharpFinite-volume IBM for membrane distillation with Robin BCsDesalination2020
Mohammadi and Nassab, 2021 [158]FVMSharpFVM–IBM for radiative heat transfer in irregular geometriesRadiative transfer2021
Mohammadi and Nassab, 2021 [159]Hybrid (LBM–FVM)SharpLBM–FVM–IBM for radiative–convective heat transferCombined heat transfer2021
Riahi et al., 2023 [160]FVMSharpDiscrete IBM for compressible flows with heat transferCompressible heat transfer2023
Ahn et al., 2023 [161]FVMSharpIBM for conjugate heat transfer with melting/solidificationPhase-change heat transfer2023
Cruz and Lamballais, 2023 [162]FDMSharpIBM for high-fidelity CHT in pipe flows with irregular geometriesConjugate heat transfer2023
Wang et al., 2022 [163]FDMSharp (direct forcing)Improved direct-forcing IBM for particle-laden flows with heat transferMultiphase heat transfer2022
Narváez et al., 2021 [164]InCompact3DSharp (dual IBM)Dual IBM for turbulent flow with conjugate heat transferTurbulent heat transfer2021
Wu et al., 2022 [165]LBMSharp (EIB)Explicit IB–RTLBFS for TFSI problems with Dirichlet BCsThermal-fluid–structure interaction2022
Zhao and Yan, 2022 [166]FEMSharp (EIBM)Enriched IBM for interface-coupled multiphysics (eg, convective heat transfer)Conjugate heat transfer2022
Ou et al., 2022 [167]DINOSharp (GCIB)Directional ghost-cell IBM for low-Mach reacting flowsReacting flows2022
Xia et al., 2014 [168]FDMSharp (ghost cell)High-order ghost-cell IBM for heat transfer; reduced grid points by ~2/3Thermal flows2014
Tao et al., 2022 [169]Other (DUGKS)Sharp (ghost cell)DUGKS–IBM for particulate flows with heat transferFluid–solid heat transfer2022
Fernandez et al., 2011 [170]Spectral MethodSharp (IBC)Grid-independent 3D heat conduction in slots with time-dependent boundariesHeat conduction2011
Shrivastava et al., 2013 [171]FVMSharp (LSIBM)Level set-based IBM for transient CFD with moving boundariesDynamic boundary flows2013
Ménez et al., 2023 [172]FVMSharp/diffuseComparison of volume penalization and IBM for compressible flows with thermal BCsCompressible thermal flows2023
Table 5. Recent advancements in IBM considering boundary treatment and corresponding fluid solvers for multiphysics applications.
Table 5. Recent advancements in IBM considering boundary treatment and corresponding fluid solvers for multiphysics applications.
ReferenceFluid SolverBoundary TreatmentSummaryApplication AreaYear
Zhang et al., 2019 [173]LBM (CLBM)DiffuseCoupled DEM–IB–CLBM framework for erosive particle impactsErosive particle flows2019
Yang et al., 2019 [174]FDM (DNS)DiffuseEulerian method for optimal perturbations in particle-laden flowsTurbulent channel flows2019
Wang et al., 2021 [175]LBMDiffusePolygonal DEM–LBM coupling with energy-conserving contact algorithmArbitrarily shaped particles2021
Zhang et al., 2021 [176]LBMDiffuseIB–LBM for coffee-ring formation in evaporating dropletsDroplet evaporation2021
Romanus et al., 2021 [177]LBMDiffuseDomain-transferring LBM–IBM for high-Re particle settlingNon-spherical particle dynamics2021
Wang et al., 2022 [178]LBMDiffuseIB–LBM for elliptical particle deposition in viscous flowsParticle deposition2022
Fukui and Kawaguchi, 2022 [179]LBMDiffuseMicroscopic particle arrangement effects on suspension rheologyNarrow channel suspensions2022
Cheng and Wachs, 2022 [180]LBM (MRT)DiffuseAdaptive octree-grid IB–LBM for particle-resolved flowsParticle-laden flows2022
Yadav and Ghosh, 2022 [181]FDMDiffuseIBM for settling of permeable/impermeable planktonic particlesBiological suspensions2022
Kawaguchi et al., 2022 [182]LBM (RLBM)DiffuseComparison of VFM and IBM for suspension flowsSuspension rheology2022
Ghosh and Panghal, 2022 [183] FDMDiffuseIBM for flexible circular particle settlingFlexible particle dynamics2022
Chang et al., 2023 [184]Other (PDDO)DiffuseHybrid peridynamic–Eulerian–IBM for FSIFluid–structure interaction2023
Panghal and Ghosh, 2023 [185]FDMDiffuseIBM for flexible/permeable planktonic particle settlingBiological sedimentology2023
Zhang et al., 2023 [186]LBMDiffuseIB–LBM for nanofluid droplet freezing with particle expulsionNanofluid dynamics2023
Yadav et al., 2023 [187]FDMDiffuseIBM for semi-torus-shaped permeable particle settlingPermeable particle dynamics2023
Ghosh et al., 2023 [188]LBMDiffuseStabilized IBM for light particles (density ratio ≥004)Light particle suspensions2023
Patel and Natarajan, 2018 [189]FVMDiffuseInterpolation-free IBM for multiphase flows with moving bodiesMultiphase FSI2018
Souza et al., 2022 [190]FEM (LES)DiffuseAMR–VOF–IBM for turbulent multiphase FSIIndustrial multiphase flows2022
Niu et al., 2022 [191]LBMDiffuseSimple diffuse IB scheme for curved boundaries in multiphase flowsBinary/multiphase flows2022
Wang et al., 2023 [192]FDMDiffuseEnergy-stable IBM for deformable membranes with non-uniform propertiesBiological membranes2023
Wang and Tian, 2019 [136]FDMDiffuseIBM for flapping wing FSI and acoustics at Mach 01Bio-inspired aerodynamics2019
Wang and Tian, 2020 [135]FEMDiffuseSound generation by 3D flexible flapping wings during hoveringBioacoustics2020
Ye et al., 2020 [193]FVMDiffuseDiscrete-forcing IBM for ship hydrodynamics with VOFMarine engineering2020
Bilbao, 2023 [194]FDTDDiffuse1D IBM for impedance/acoustic barriersLinear acoustics2023
Bilbao, 2023 [195]FDTDDiffuse3D IBM for irregular boundaries in wave-based acoustics3D acoustic simulations2023
Hou et al., 2023 [196]FDTDDiffuseTime-domain IBM for acoustic propagation between gas bubblesBubble acoustics2023
Jiang et al., 2022 [142]LBMDiffuse (BTDF–IBM)Parallel FR–DNS for settling suspensions with lubrication modelLarge-scale suspensions2022
Nangia et al., 2019 [197]FDMDiffuse (DLM–IBM)DLM–IBM for high-density ratio WSI with AMRWave-structure interaction2019
Cheng et al., 2021 [198]FDMDiffuseSemi-implicit IBM for viscous flow-induced soundComputational aeroacoustics2021
Zeng et al., 2022 [199]Godunov schemeDiffuse (DLM–IBM)Adaptive DLM–IBM with subcycling for single/multiphase FSIAdaptive FSI2022
Yan et al., 2022 [200]LBM (MLBFS)DiffuseIB–MLBFS with flux correction for multiphase FSIMultiphase FSI2022
Hori et al., 2022 [201]SACDiffuse (implicit)Eulerian-based IBM with implicit lubrication for particle suspensionsDense suspensions2022
Liu et al., 2017 [202]DNS (FDM)Diffuse (point-particle)Combined DNS, IBM, and DPM for sediment particle distribution in turbulent boundary layersSediment transport2017
Wang et al., 2017 [203]FDM (WENO)DiffuseCompressible multiphase FSI with ANCF structural solverExplosive/impact dynamics2017
Cheng et al., 2017 [204]FDMDiffuse (smoothed IB)Compressible IBM for flow-induced noise using influence matricesAeroacoustics2017
Zhang et al., 2013 [205]FDMDiffuse (VOF-IBM)Two-phase IB–VOF model for ocean engineering problemsFree-surface flows2013
Bürchner et al., 2023 [206]FEM (FCM)Diffuse (γ-scaling)FWI with γ-scaling IBM for crack detection in NDTNon-destructive testing2023
Fazli et al., 2023 [207]FDMDiffuseIBM for yield-pseudoplastic particulate flowsNon-Newtonian suspensions2023
Chéron et al., 2023 [208]FVMHybrid (sharp–diffuse)Hybrid IBM for dense particle-laden flows with non-symmetrical operatorsDense suspensions2023
Ido et al., 2017 [209]LBMSharpHybrid LBM–IBM–DEM for magnetic particle microstructures in MR fluidsMagnetorheological fluids2017
Fukui et al., 2018 [210]LBMSharpTwo-way coupling for particle rotation effects on suspension rheologySuspension rheology2018
Barbeau et al., 2022 [211]FEMSharpHigh-order FEM–IBM for flow around sphere packingsFixed-bed reactors2022
Wu and Chen, 2022 [212]LBMSharp3D IB–LBM for droplet-particle coating processesSpray coating2022
Duprez et al., 2023 [213]FEM (ϕ-FEM)Sharpϕ-FEM for Stokes flow around particles with optimal convergenceCreeping particle flows2023
Farooq et al., 2025 [214]QUICKSharpRadial-basis function for incompressible flowBio-inspired moving bodies2025
Zhao et al., 2021 [215]FDM (DNS)SharpSharp-interface IBM–APE for flow-induced noise predictionCylinder array acoustics2021
Gorges et al., 2024 [216]FVMSharp (blocked off)Comparison of smooth/blocked-off IBM for fixed-bed reactorsPacked bed reactors2024
Zhou and Balachandar, 2021 [217]FDMSharp (direct forcing)Analysis of spatiotemporal resolution in IBM with direct forcingRigid particulate flows2021
Xie et al., 2020 [218]FDM (DRP)Sharp (ghost cell)Cartesian grid method for acoustic scattering using ghost-cell IBMAcoustic scattering2020
Zhao et al., 2022 [219]FDTDSharp (ghost node)FDTD–IBM for underwater acoustic scatteringUnderwater acoustics2022
Isoz et al., 2022 [220]FVMSharp (hybrid)Hybrid fictitious domain–IBM for irregular particle flowsIndustrial particle flows2022
Qin et al., 2022 [221]FDMSharpHybrid IBM for particle–complex boundary interactionsParticulate flows2022
Yu et al., 2023 [127]FDM/WENOSharp (IB–GHPC)Two-phase wave tank with GHPC pressure solverCoastal engineering2023
Giahi and Bergstrom, 2023 [222]FVMSharp (IBS)Validation of IBS–IBM for fixed/moving bodiesGeneral FSI2023
Zeng et al., 2021 [223]FVMSharp (multi-sphere)Lagrangian IBM for proppant transport in hydraulic fracturesOil/gas industry2021
Caunt et al., 2023 [224]FDMSharp (Taylor series)IBM for seismic/acoustic wave propagation over irregular terrainGeophysics2023
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

Kanchan, M.; Kumar, A.M.; Arun, P.A.H.; Powar, O.; Mehar, K.; Mangalore, P. Advances in Flow–Structure Interaction and Multiphysics Applications: An Immersed Boundary Perspective. Fluids 2025, 10, 217. https://doi.org/10.3390/fluids10080217

AMA Style

Kanchan M, Kumar AM, Arun PAH, Powar O, Mehar K, Mangalore P. Advances in Flow–Structure Interaction and Multiphysics Applications: An Immersed Boundary Perspective. Fluids. 2025; 10(8):217. https://doi.org/10.3390/fluids10080217

Chicago/Turabian Style

Kanchan, Mithun, Anwak Manoj Kumar, Pedapudi Anantha Hari Arun, Omkar Powar, Kulmani Mehar, and Poornesh Mangalore. 2025. "Advances in Flow–Structure Interaction and Multiphysics Applications: An Immersed Boundary Perspective" Fluids 10, no. 8: 217. https://doi.org/10.3390/fluids10080217

APA Style

Kanchan, M., Kumar, A. M., Arun, P. A. H., Powar, O., Mehar, K., & Mangalore, P. (2025). Advances in Flow–Structure Interaction and Multiphysics Applications: An Immersed Boundary Perspective. Fluids, 10(8), 217. https://doi.org/10.3390/fluids10080217

Article Metrics

Back to TopTop