Fact-checked by Grok 2 weeks ago

Finite volume method

The finite volume method (FVM) is a numerical technique for solving partial differential equations (PDEs), particularly those derived from physical conservation laws such as , , and balances in and . It operates by partitioning the computational domain into a finite number of non-overlapping s, applying the integral form of the governing conservation equations to each volume, and approximating surface fluxes to ensure local and global conservation of physical quantities. This approach yields a of algebraic equations solved iteratively, with the represented as piecewise constants or linear interpolations within each . The method emerged in the 1970s as part of broader advancements in (CFD), alongside developments in upwind and finite element methods, enabling the of complex flows in applications. It gained prominence through the work of Suhas V. Patankar, whose 1980 book Numerical Heat Transfer and Fluid Flow systematized FVM for practical use in convection-diffusion problems, introducing key discretization strategies like the hybrid scheme for handling convective fluxes. Earlier roots trace to box methods and integral formulations in the 1960s and 1970s for elliptic PDEs, but Patankar's contributions emphasized its robustness for structured grids and pressure-velocity coupling in incompressible flows via algorithms like . At its core, FVM leverages the to convert volume integrals of PDEs into surface integrals of fluxes, facilitating straightforward implementation on arbitrary geometries without requiring coordinate transformations. Fluxes at interfaces are approximated using upwind biasing for in problems or central differencing for elliptic ones, with time often via explicit or implicit schemes. Compared to methods, which rely on differential forms and structured grids, FVM offers superior conservation properties and adaptability to complex boundaries; relative to finite element methods, it avoids variational principles, directly enforcing balance equations for enhanced physical fidelity. FVM's strengths lie in its inherent conservation, second-order accuracy on refined meshes, and versatility across PDE types—from hyperbolic conservation laws in shock-capturing to elliptic equations in porous media flow—making it a cornerstone of modern CFD software like ANSYS Fluent and . Applications span (turbulent airfoil simulations), (reactor thermal hydraulics), and (pollutant dispersion), with extensions to multiphase flows, reactive systems, and even via finite volume formulations. Despite challenges like flux limiter selection for high-speed flows, ongoing research as of 2025 focuses on high-order schemes and accelerations to broaden its scope.

Overview

Definition and core principles

The finite volume method (FVM) is a numerical technique for solving partial differential equations (PDEs) that arise from laws, such as those governing mass, momentum, and energy in and . It divides the computational into a finite number of non-overlapping control volumes, or , and enforces local by integrating the governing equations over each . This approach ensures that the total amount of conserved quantities is preserved across the , making FVM particularly robust for problems involving discontinuities or shocks. The core principle of FVM is flux conservation, which states that the rate of change of a within a equals the net through its boundaries minus any internal sources or sinks. In semi-discrete form, this is expressed as \frac{d}{dt} \int_V u \, dV = -\oint_S \mathbf{F} \cdot \mathbf{n} \, dS + \int_V s \, dV, where V is the , S its boundary surface, u the conserved variable (e.g., or ), \mathbf{F} the vector, \mathbf{n} the outward unit normal, and s the source term. This integral formulation approximates the PDEs by balancing fluxes at cell interfaces, promoting numerical stability and accuracy in conserving physical quantities. FVM implementations typically employ a cell-centered approach, where primary variables are stored at cell centroids and fluxes are computed at faces, though vertex-centered variants store values at mesh vertices and average them for cell integrals; the cell-centered method remains the standard due to its simplicity and direct enforcement of conservation. To illustrate in one dimension, consider the scalar advection equation \partial_t u + a \partial_x u = 0 (with constant speed a > 0) over a domain divided into cells [x_{i-1/2}, x_{i+1/2}]. The FVM semi-discrete approximation for cell i becomes \frac{d}{dt} ( \bar{u}_i \Delta x_i ) = - (F_{i+1/2} - F_{i-1/2} ), where \bar{u}_i is the cell average and F_{i+1/2} approximates the flux at the interface, such as via upwinding: F_{i+1/2} = a \bar{u}_i. This discretizes the integral form while maintaining conservation. Unlike finite difference methods, which are collocation-based, or finite element methods, which use variational principles, FVM is inherently integral-based and excels in unstructured grids for conservation-dominated problems.

Historical development

The finite volume method (FVM) emerged from foundational work on numerical solutions to partial differential equations in fluid dynamics during the mid-20th century, building on finite difference techniques. Foundational contributions include the 1928 analysis by Richard Courant, Kurt Otto Friedrichs, and Hans Lewy of finite difference methods for hyperbolic equations, introducing the Courant-Friedrichs-Lewy (CFL) stability condition, which ensures that information propagation in numerical schemes aligns with physical wave speeds and remains essential for FVM stability. Concurrently, John von Neumann and Robert D. Richtmyer advanced shock-capturing capabilities in 1950 by developing an artificial viscosity method to handle discontinuities in compressible flows without explicit shock fitting, marking an early precursor to conservative finite volume discretizations. Sergei K. Godunov further solidified these foundations in 1959 with the first explicit finite volume scheme for gas dynamics equations, employing exact Riemann solvers to compute cell-interface fluxes while preserving conservation properties. The 1970s saw the formalization and practical expansion of FVM, particularly for hyperbolic conservation laws and shock-capturing applications. Robert D. Richtmyer, along with Kenneth W. Morton, contributed to this phase through their 1967 monograph on difference methods for initial-value problems, which analyzed and accuracy in schemes relevant to FVM precursors for fluid flows. A pivotal milestone occurred with the development of FVM for and incompressible by V. Patankar, in collaboration with Brian Spalding's group at , extended integral conservation formulations to structured grids. This led to the creation of the TEACH (Teaching) code around 1974, an early computational tool for solving coupled convection-diffusion equations in engineering applications. Patankar's influential 1980 book, Numerical and Fluid Flow, systematized the approach, emphasizing staggered grids and the pressure-velocity coupling algorithm, and became a cornerstone reference for FVM implementation in thermal and fluid sciences. During the 1980s and 1990s, FVM expanded significantly into high-resolution methods for compressible flows, with notable contributions from , , and others building on Godunov's framework. and introduced flux-corrected transport (FCT) in , a monotonicity-preserving technique that combined low-order with anti-diffusive corrections to achieve sharp shock resolution without oscillations, influencing subsequent Godunov-type finite volume schemes for multidimensional simulations. These methods gained traction in the 1980s for handling complex compressible phenomena, as seen in extensions like the piecewise parabolic method (PPM) by Paul Woodward and Phillip Colella in 1984, which improved accuracy on unstructured grids. Post-2000 advancements have focused on higher-order accuracy, adaptive techniques, and applicability to unstructured meshes, enhancing FVM's versatility in large-scale simulations. Nishikawa has been a key figure in this era, developing hyperbolic reformulations for terms to enable efficient high-order finite volume schemes, as detailed in his work on hyperbolic systems for time-dependent . His contributions, including third-order accurate, second-derivative-free discretizations by 2023, have advanced shock-capturing and viscous flow predictions on complex geometries. This evolution reflects a timeline of major publications, from Godunov's 1959 paper and Richtmyer-Morton's 1967 book to Patankar's 1980 text and modern high-order extensions, underscoring FVM's shift from basic conservation enforcement to sophisticated multiphysics solvers.

Mathematical formulation

Conservation laws in integral form

The finite volume method is grounded in the integral formulation of laws, which express the balance of a within a . The starting point is the general (PDE) for a u(\mathbf{x}, t) in three dimensions: \frac{\partial u}{\partial t} + \nabla \cdot \mathbf{F}(u) = s(u), where \mathbf{F}(u) is the representing the of u across surfaces, and s(u) denotes source terms accounting for production or consumption within the volume. This assumes sufficient smoothness of u and \mathbf{F}. To obtain the integral form, integrate the PDE over an arbitrary fixed V bounded by surface S, and apply the to the flux term: \frac{d}{dt} \int_V u \, dV + \oint_S \mathbf{F} \cdot \mathbf{n} \, dS = \int_V s \, dV, where \mathbf{n} is the outward unit normal to S. This equation states that the rate of change of the total amount of u in V equals the net out of S plus the integrated sources inside V. The integral form is more general, as it holds even for discontinuous solutions without requiring differentiability. Specific cases illustrate the flux structure. For scalar advection, the flux is \mathbf{F} = a u, where a is the advection velocity, modeling transport without diffusion or sources (assuming s = 0). In diffusion problems, \mathbf{F} = -D \nabla u, with D > 0 the diffusion coefficient, capturing spreading due to gradients. The Navier-Stokes equations for compressible fluids extend this to a system of conservation laws for mass, momentum, and energy, with total fluxes including convective, pressure, and diffusive terms, such as \mathbf{F} = \rho \mathbf{u} \otimes \mathbf{u} + p \mathbf{I} - \boldsymbol{\tau} for momentum (where \rho is density, \mathbf{u} velocity, p pressure, \mathbf{I} identity, and \boldsymbol{\tau} viscous stress tensor) and source terms for body forces such as gravity. Conservation laws are classified by PDE type, influencing finite volume method (FVM) design and stability. Hyperbolic PDEs, like pure advection (\frac{\partial u}{\partial t} + a \frac{\partial u}{\partial x} = 0), feature finite wave speeds and possible discontinuities, requiring upwind schemes in FVM for stability via the CFL condition. Parabolic PDEs, such as the diffusion equation (\frac{\partial u}{\partial t} = D \frac{\partial^2 u}{\partial x^2}), involve infinite propagation speeds and smoothing, where FVM uses central differencing but needs careful treatment for mixed advection-diffusion to avoid oscillations. Elliptic PDEs, like steady-state Poisson equations, lack time derivatives and are less common in time-dependent FVM applications, often solved iteratively for boundary-value problems. These classifications guide FVM stability: hyperbolic systems demand monotonicity-preserving methods, while parabolic ones benefit from positivity enforcement. For problems with discontinuities, such as in hyperbolic laws, classical solutions fail, necessitating weak solutions that satisfy the integral form in a distributional sense. Weak solutions allow jumps in u while conserving the total quantity across interfaces, as verified by the Rankine-Hugoniot condition for shock speeds. In FVM, approximating weak solutions ensures physical accuracy, particularly for nonlinear systems where multiple weak solutions exist, resolved via conditions to select the physically admissible one.

Flux and source terms

In the integral form of conservation laws, the flux terms represent the net transport of conserved quantities across the boundaries of a control volume, while source terms account for internal generation or consumption within the volume itself. These components arise naturally from the applied to the differential form of the equations, ensuring that the finite volume method preserves the physical principles locally. Flux terms are typically decomposed into convective and diffusive contributions. The convective flux, \mathbf{F}_c = u \mathbf{v}, captures the transport of a scalar quantity u (such as mass fraction or momentum) by the bulk velocity \mathbf{v}, as seen in the momentum equation where it appears as \rho \mathbf{v} \otimes \mathbf{v} for the advective transport of momentum density \rho \mathbf{v}. In contrast, the diffusive flux, \mathbf{F}_d = -\Gamma \nabla u, models transport due to gradients, driven by molecular diffusion or viscosity; for the energy equation, the diffusive flux includes the conductive heat flux -\kappa \nabla T, with viscous dissipation treated as a source term, where \Gamma is the diffusion coefficient and \kappa the thermal conductivity. This separation allows tailored numerical treatments, with convective terms often requiring upwind stabilization to handle hyperbolic behavior, while diffusive terms benefit from central differencing for elliptic accuracy. Source terms, denoted as \mathbf{S}, represent volumetric effects integrated over the control volume and include phenomena like chemical reaction rates or external forces. For reacting flows, sources may involve production rates such as \dot{\omega} = k c^n for species concentration c, where k is the and n the , directly added to the species equation. Gravitational sources appear as \rho \mathbf{g} in the momentum equation, contributing a body force per unit volume that is simply the times the vector \mathbf{g}. These terms are evaluated as volume averages, \bar{S} = \frac{1}{V} \int_V S \, dV, to maintain local without surface fluxes. In systems of nonlinear conservation laws, such as the Euler equations for inviscid , the flux becomes a nonlinear function of the conserved variables. The one-dimensional flux vector is given by \mathbf{F} = \begin{pmatrix} \rho v \\ \rho v^2 + p \\ \rho v E + p v \end{pmatrix}, where \rho is , v , p , and E total per unit ; this form ensures the method captures shocks and rarefactions accurately through characteristic decomposition. The nonlinearity introduces coupling between equations, necessitating flux Jacobians for stability analysis in Godunov-type schemes. Non-conservative products and geometric sources arise in formulations involving variable coefficients or non-Cartesian grids, particularly in . Non-conservative terms, like those in the with (h \frac{\partial z_b}{\partial x}), require path-conservative schemes to define consistent integrals along characteristics, avoiding spurious oscillations. In curvilinear systems, geometric sources emerge from coordinate transformations, such as centrifugal terms -\rho \frac{u_\theta^2}{r} in cylindrical coordinates for the radial momentum equation, which must be included to maintain and prevent errors in unstructured meshes. These are often treated as volume sources to preserve the conservative flux differencing. The proper formulation of fluxes and sources ensures , meaning the numerical solution remains unchanged under constant-velocity frame shifts, which is crucial for frame-independent conservation in . Conservative of fluxes guarantees this property for systems, as non-conservative handling can introduce frame-dependent errors in shock positions or . This invariance is verified by transforming the equations under Galilean boosts and confirming identical discrete forms.

Discretization process

Control volume setup

The finite volume method begins by partitioning the physical domain into a set of non-overlapping , which serve as the fundamental elements for discretizing the governing equations. This domain decomposition is crucial for applying the integral form of laws to each volume, ensuring that the method inherently satisfies local and global conservation properties. The choice of partitioning strategy significantly influences the method's accuracy, efficiency, and applicability to complex geometries. Domain decomposition can be structured or unstructured, each with distinct characteristics suited to different problem types. Structured grids, often Cartesian or curvilinear with hexahedral cells, arrange control volumes in a regular indexed by (i,j,k), facilitating efficient and implicit of neighboring volumes. These grids excel in problems with simple geometries, offering advantages in computational speed and ease of for higher-order schemes, but they lack flexibility for irregular boundaries, potentially requiring multi-block topologies that complicate . In contrast, unstructured grids employ polyhedral or tetrahedral cells without predefined ordering, relying on explicit lists; this approach provides superior adaptability to complex shapes, enabling easier of and refinement, though it demands more memory for and increases bookkeeping complexity due to irregular neighbor relations. Hybrid strategies combine structured regions for boundary layers with unstructured interiors to balance efficiency and flexibility. Control volumes in the finite volume method are defined either as cells in cell-centered approaches or as meshes in vertex-centered (node-centered) formulations. In cell-centered methods, the primary mesh cells themselves act as control volumes, with solution variables stored at cell centroids, leading to compact stencils with bandwidth equal to the number of neighbors plus one, which promotes efficient matrix solvers but imposes stricter requirements on mesh to minimize errors on non-uniform grids. Vertex-centered methods construct control volumes around mesh vertices by forming median- partitions—connecting cell centers to face midpoints and edge midpoints—resulting in more flexible handling of unstructured topologies and potentially more accurate gradients via least-squares , albeit at the cost of larger stencils and higher storage needs. Both approaches approximate the solution as uniform within each control volume, with fluxes evaluated at faces shared between adjacent volumes. Once defined, the geometric properties of control volumes, such as volume and face areas, must be computed accurately, particularly on unstructured meshes where cells may be irregular. For a control volume i in d-dimensions (typically d=2 or $3), the volume V_i is calculated using the divergence theorem applied to the position vector: V_i = \frac{1}{d} \sum_f (\mathbf{r}_f \cdot \mathbf{A}_f), where the sum is over all faces f bounding the volume, \mathbf{r}_f is the position vector of the face centroid, and \mathbf{A}_f is the outward-facing area vector with magnitude equal to the face area and direction normal to the face. Face areas A_f are determined by projecting the face geometry onto coordinate planes or using vector cross-products for polygonal faces, ensuring the normal points consistently outward for conservation. In vertex-centered setups on triangular or tetrahedral primal meshes, the control volume is the union of sub-regions around the vertex (e.g., connecting barycenters of adjacent cells to edge midpoints), with V_i as the sum of these sub-volumes' areas in 2D or volumes in 3D. Accurate computation of these properties is essential, as errors can propagate into flux approximations, though second-order spatial accuracy is preserved on sufficiently smooth grids. Grid generation for control volumes requires careful attention to quality metrics like , , and , as these directly impact numerical accuracy and stability. measures the alignment of face normals with the line connecting centroids; high reduces truncation errors in computations, particularly for diffusive terms, while non-orthogonal grids necessitate corrections like over-relaxed approaches to maintain second-order accuracy. , quantifying deviation from ideal shapes (e.g., equilateral triangles or cubes), amplifies interpolation errors and can degrade , with severe limiting the allowable and time-step size. , the ratio of maximum to minimum dimensions, should ideally approach unity for isotropic flows but can be high (e.g., >100) in boundary layers to resolve gradients efficiently; excessive ratios, however, couple with to introduce numerical or oscillations. Tools for grid generation, such as advancing front or for unstructured meshes, aim to minimize these metrics while conforming to boundaries. To enhance resolution in regions of high gradients, such as shocks or boundary layers, adaptive mesh refinement (AMR) introduces local h-refinement, which subdivides control volumes without altering the overall . In h-refinement, coarser s are bisected (e.g., a hexahedral into eight smaller ones in , halving the edge length), guided by error estimators like undivided differences in flow variables or residual thresholds, ensuring hanging-node conformity by limiting level differences between adjacent cells to one. This technique efficiently allocates computational resources, achieving grid convergence with reduced cost—for instance, refining to levels where size is h/16 can resolve features while using only 5% of the runtime of uniform refinement—though it requires careful balancing at non-conformal interfaces to preserve .

Spatial reconstruction and flux computation

Spatial reconstruction in the finite volume method approximates the conserved variables within each to determine values at cell interfaces, allowing for higher-order spatial accuracy beyond the basic cell-averaged representation. The zeroth-order reconstruction assumes a constant profile within the cell, setting the left and right interface states equal to the cell average u_i, which yields a accurate scheme suitable for initial implementations but prone to excessive numerical diffusion. To achieve second-order accuracy, linear (piecewise linear) reconstruction extrapolates the solution using cell-centered gradients, as in the Monotonic Upstream-centered Scheme for Conservation Laws (MUSCL), which computes interface states u_{i+1/2}^L = u_i + \frac{1}{2} \phi(r) (u_i - u_{i-1}) and similarly for the right state, where \phi(r) is a function based on the ratio r of consecutive gradients to suppress oscillations near discontinuities. Higher-order reconstructions, such as weighted essentially non-oscillatory () schemes, adaptively combine stencils to maintain smoothness in smooth regions while avoiding Gibbs phenomena across shocks, achieving up to fifth-order accuracy by weighting candidate polynomials based on their smoothness indicators. Flux computation evaluates the numerical flux F_{i+1/2} at cell interfaces to update cell averages, typically as a of the reconstructed left (u_L) and right (u_R) states: F_{i+1/2} = f(u_L, u_R), resolving the local Riemann problem arising from the discontinuity at the interface. For hyperbolic conservation laws, central differencing averages fluxes from both sides but can introduce instabilities; upwind schemes, such as Godunov's method, instead solve the exact Riemann problem with piecewise constant states to select physically relevant waves, ensuring monotonicity and stability. Approximate Riemann solvers improve efficiency: Roe's approximates the for wave propagation, while the HLLC solver captures contact discontinuities by including two intermediate waves between the fastest and slowest signals, reducing dissipation compared to the simpler HLL variant. Limiter functions, such as the minmod limiter \phi(r) = \max(0, \min(1, r)) or van Leer's \phi(r) = \frac{2r}{1+r} (for r > 0), are applied during to enforce (TVD) properties, preventing new extrema and ensuring the scheme remains monotonic for scalar equations, as proven in the framework of high-resolution schemes. These properties guarantee that the of the solution does not increase over time steps, promoting and to entropy-satisfying weak solutions even on coarse grids. For problems involving diffusive terms, such as in the Navier-Stokes equations, the viscous fluxes are discretized separately using reconstructed gradients at interfaces, often via central differencing for symmetry: the diffusive flux component approximates \mathbf{\Gamma}_{i+1/2} \approx \frac{\partial \mathbf{u}}{\partial \mathbf{x}} \big|_{i+1/2}, where \mathbf{\Gamma} represents the tensor. Explicit treatment of these terms follows the convective fluxes but restricts time steps due to stiffness from small diffusivities; implicit discretization, solving a for the contribution, allows larger time steps while maintaining second-order accuracy and unconditional for pure problems.

Numerical implementation

Time integration schemes

After the spatial discretization in the finite volume method, the governing laws are reduced to a semi-discrete system of ordinary differential equations (ODEs) for the cell-averaged conserved variables u_i: \frac{d u_i}{dt} = R(u_i), where R(u_i) denotes the spatial , which encapsulates the net flux balance across cell interfaces and any source terms. This ODE system must then be in time to obtain the fully discrete solution, transforming the partial differential equations into a sequence of time levels. The choice of time influences the accuracy, , and computational , particularly in handling or stiff systems common in (CFD). Explicit time integration schemes evaluate the residual R using known values from previous time levels to directly compute the update, making them computationally inexpensive per step but subject to stability restrictions. The forward Euler method, a first-order explicit scheme, updates the solution as u_i^{n+1} = u_i^n + \Delta t \, R(u_i^n), where \Delta t is the time step and superscript n denotes the time level; it is simple but prone to excessive numerical diffusion and limited by stringent stability criteria. Higher-order explicit methods, such as Runge-Kutta (RK) schemes, mitigate these issues by performing multiple internal stages per time step to achieve greater accuracy. A seminal application of explicit RK time stepping in finite volume methods for solving the Euler equations demonstrated its effectiveness in capturing shocks and complex flows on unstructured grids, with second- and third-order variants providing improved resolution over Euler. For instance, low-storage third-order RK schemes, designed for efficiency in large-scale CFD simulations, require only a few registers per variable and maintain third-order accuracy with a local truncation error of O(\Delta t^4), while offering good stability for mildly nonlinear problems. These multistage RK methods generally attain an order of accuracy p through optimized stage coefficients, and error estimation can be incorporated via embedded lower-order companions to adaptively control \Delta t. However, explicit schemes' stability is governed by the Courant-Friedrichs-Lewy (CFL) condition, which requires \Delta t \leq \frac{\Delta x}{\max |a|} for a one-dimensional advection problem with characteristic speed a, ensuring the numerical domain of dependence encompasses the physical one; violations lead to instability, often limiting \Delta t to a fraction of the spatial grid size \Delta x. Implicit time integration schemes address in problems dominated by or low-speed flows by incorporating the at the future time level, typically requiring the solution of a linear or at each step. The , a implicit scheme, is formulated as u_i^{n+1} - u_i^n = \Delta t \, R(u_i^{n+1}), offering unconditional for linear diffusive systems and robustness against oscillations in stiff regimes, though at the cost of reduced formal accuracy. The Crank-Nicolson scheme improves upon this by averaging the between time levels: u_i^{n+1} - u_i^n = \frac{\Delta t}{2} \left[ R(u_i^n) + R(u_i^{n+1}) \right], achieving second-order accuracy while preserving stability for parabolic equations; it is particularly effective in finite volume discretizations of the Navier-Stokes equations where diffusion terms stiffen the system. These implicit methods allow larger time steps than explicit ones, beneficial for steady-state convergence or low-Mach-number flows, but their computational overhead from matrix inversions scales poorly for three-dimensional simulations without preconditioning. For unsteady simulations requiring accurate resolution of physical time scales alongside efficient inner iterations, dual-time stepping methods decouple the physical time advancement from pseudo-time iterations to reach a steady state within each physical step. This approach formulates the semi-discrete system with an additional pseudo-time derivative: \frac{\partial u}{\partial \tau} + \frac{\partial u}{\partial t} = R(u), where \tau is the pseudo-time, and iterates explicitly or implicitly in \tau until convergence for fixed physical time t, enabling larger effective time steps in applications like turbomachinery where periodic unsteadiness demands high temporal fidelity. Dual-time stepping with implicit RK inner solvers has been shown to enhance efficiency for fully unsteady flows, maintaining high-order accuracy while mitigating stiffness from spatial residuals. In multistage contexts, such as RK-based dual-time methods, the order of accuracy is preserved across physical steps, with error estimation derived from pseudo-time residuals to ensure convergence criteria are met without over-resolving transients.

Boundary condition handling

In the finite volume method (FVM), boundary conditions are incorporated to ensure the laws are satisfied across the entire computational , including at its edges, by appropriately defining the fluxes through boundary faces. This maintains the integral form of the governing equations while reflecting physical constraints such as fixed values, specified gradients, or cyclic behavior. Accurate handling of boundaries is essential for and to prevent unphysical reflections or imbalances in the solution. The primary types of boundary conditions in FVM include Dirichlet, , Robin, and periodic conditions. Dirichlet conditions prescribe a fixed value for the dependent variable at the boundary, such as a specified or on a . conditions specify the or of the variable, for instance, imposing zero on an insulated surface or a prescribed at an . Robin conditions combine elements of Dirichlet and types through a involving both the variable value and its , commonly used for convective where the boundary depends on an external . Periodic conditions enforce continuity across opposite boundaries by equating variables and fluxes, simulating repeating domains like channel flows. For structured grids, boundary conditions are often implemented using ghost cells or image points, which extend the domain with fictitious cells adjacent to the physical boundary to facilitate flux computations without altering the internal stencil. In this approach, values in the ghost cells are set to satisfy the desired condition; for example, under a Dirichlet condition, the ghost cell value mirrors the prescribed boundary value, while for Neumann, it uses extrapolation to enforce zero gradient. This method simplifies the application of interface fluxes from spatial reconstruction at boundaries by treating boundary faces as internal ones. Flux adjustments at boundaries modify the net flux into control volumes near the edge to enforce specific physical behaviors. For a no-slip , the tangential is extrapolated to zero at the while the normal component is set to zero, ensuring adherence to the wall without slip. At an , a specified or profile is imposed by adjusting the incoming flux, often treating it as a source term in the discretized to maintain mass conservation. These adjustments are computed using the same reconstruction schemes as internal faces but with boundary-specific extrapolations or constraints. In unstructured grids, boundary treatment relies on identifying boundary faces shared by a single control volume, where fluxes are evaluated directly using half-volumes or virtual cells to approximate the boundary contribution. Boundary faces are defined as the intersection of a cell with the domain edge, and fluxes through them incorporate the boundary data via owner-neighbor connectivity, often with corrections for non-orthogonality. Virtual cells, akin to ghost cells, may be introduced outside the domain for complex boundaries, setting values to reverse normal velocities or match prescribed conditions while preserving tangential components for viscous flows. This approach ensures flux balance without requiring a full extension of the grid. Special cases like far-field non-reflecting conditions are crucial for open-domain simulations, such as external , to allow outgoing waves to exit without spurious reflections. These are typically implemented using variables derived from the linearized Euler equations, decomposing the flow into wave components and absorbing only the outgoing ones at the , often with a decay term to promote . This method, based on long-wave perturbations around a steady far-field , enhances in finite volume schemes by reducing reflections compared to simple .

Advantages and limitations

Key strengths

The finite volume method (FVM) ensures intrinsic conservation of quantities such as , , and , both globally across the entire and locally within each , by directly discretizing the integral form of laws. This property arises from the method's formulation, where fluxes through cell interfaces are balanced to mimic the physical principles, making it particularly suitable for problems involving shocks, discontinuities, and wave propagation in hyperbolic systems. FVM offers significant flexibility in handling complex geometries through its compatibility with unstructured, , or adaptive meshes, unlike methods that typically require regular grids. Control volumes can be polyhedral or of varying shapes and sizes, allowing accurate representation of irregular boundaries and domains in applications without excessive computational overhead. The method demonstrates robustness for nonlinear problems, especially those dominated by , due to its natural incorporation of upwind schemes that stabilize solutions by accounting for the direction of information propagation. Upwinding prevents oscillations near discontinuities and enhances monotonicity, providing reliable performance for systems like the Euler equations in . Implementation of complex physics is straightforward in FVM, as source terms for multiphase flows, reactions, or porous media can be readily integrated into the balances without altering the core framework. This modularity facilitates extensions to coupled phenomena, such as with phase changes, maintaining the method's conservative structure. In contexts, FVM achieves proven second-order accuracy on smooth solutions through techniques like linear and limiters, while inherently preserving positivity for quantities like or in problems. This balance of accuracy and has made it a standard in simulations.

Common challenges and resolutions

Low-order finite volume schemes often suffer from excessive numerical and , which smear sharp gradients and reduce solution accuracy in problems involving discontinuities or thin layers. This issue arises primarily from the upwind , which introduces artificial to stabilize the scheme but at the cost of resolving fine-scale features. To address this, higher-order reconstruction techniques, such as the Monotonic Upstream-centered Scheme for Conservation Laws (MUSCL), employ slope limiters like the minmod or van Leer limiter to achieve second-order accuracy while preventing oscillations near discontinuities. These limiters blend low-order monotonic profiles with higher-order extrapolations based on local smoothness indicators, effectively reducing without compromising stability. Further improvements come from weighted essentially non-oscillatory () reconstructions, which adaptively weight stencils to minimize in smooth regions and capture shocks sharply. On unstructured grids, the finite volume method incurs high computational costs due to irregular connectivity, leading to larger sizes and more complex computations compared to structured meshes. This is exacerbated in three-dimensional simulations where assembly and inversion become bottlenecks for implicit solvers. Parallelization strategies, such as domain decomposition with message-passing interfaces, distribute workloads across processors to scale efficiently for large-scale problems. Additionally, efficient iterative solvers like algebraic multigrid () accelerate convergence by hierarchically coarsening the grid and smoothing errors at multiple levels, achieving a nearly optimal total work complexity of O(N), with the number of iterations often independent of N or growing logarithmically for N unknowns on unstructured meshes. Capturing shocks without excessive smearing poses a challenge in systems, as standard Riemann solvers may produce phenomena or entropy-violating expansions across points. Low-dissipation schemes tend to under-resolve shocks, leading to oscillations, while overly dissipative ones blur discontinuities. Shock detectors, such as those based on or jumps exceeding a , identify troubled cells and activate localized limiting. Entropy fixes in approximate Riemann solvers, like the Harten-Lax-van Leer (HLLE) or solver with added dissipation terms, enforce the condition by splitting small eigenvalues or introducing artificial near rarefactions, ensuring physical admissibility. In multiphysics simulations involving disparate timescales, such as reacting flows or low-Mach-number , the finite volume method encounters stiffness from source terms or , severely restricting explicit time steps. Hybrid explicit-implicit (IMEX) schemes treat stiff components implicitly while advancing non-stiff terms explicitly, allowing larger global time steps without resolving fast transients fully. Preconditioning techniques, such as artificial or dual-time stepping, scale the system to equalize eigenvalues, improving convergence of iterative solvers for incompressible or multiphase regimes. Highly skewed grids, common in complex geometries, degrade accuracy in finite volume methods by introducing errors in gradient reconstruction and flux evaluation due to non-orthogonal faces and distorted control volumes. Skewness amplifies truncation errors, particularly in convection-dominated flows, leading to unphysical solutions or divergence. Orthogonal quality metrics, defined as the cosine of the angle between face normals and cell centroid vectors (ideally close to 1), quantify and guide improvements. Grid optimization via Laplacian or elliptic smoothing algorithms repositions nodes to minimize skewness while preserving boundary conformity, enhancing overall solution fidelity without remeshing. Recent advances as of 2025 address ongoing computational challenges in high-order three-dimensional FVM applications, particularly high costs and the need to balance accuracy with stability in multiphase and unstable flows. Integration of , such as deep for initial solution predictions, and (HPC) enable efficient handling of larger models. Hybrid methods combining FVM with approaches or (PINNs) further improve performance while requiring specialized skills to minimize numerical errors.

Applications

Computational fluid dynamics

The finite volume method (FVM) serves as a cornerstone for solving the Navier-Stokes equations in (CFD), particularly for simulating fluid flows by integrating conservation laws over control volumes. For incompressible flows, FVM discretizes the and equations, often employing pressure-velocity coupling techniques such as the (Semi-Implicit Method for Pressure-Linked Equations) algorithm, which iteratively corrects and fields to enforce mass conservation. This approach is widely used in laminar and low-Mach-number simulations, ensuring robust handling of viscous effects without explicit variations. In contrast, for compressible flows, FVM incorporates fluctuations and equations, typically using flux-limited schemes to capture shocks and expansions accurately, as seen in high-speed applications. Turbulence modeling in FVM-based CFD predominantly relies on Reynolds-Averaged Navier-Stokes (RANS) formulations, where the method solves time-averaged equations augmented with closure models like the k-ε model to approximate turbulent stresses via eddy viscosity. The k-ε model, which transports (k) and its dissipation rate (ε), is integrated into the FVM framework on structured or unstructured grids, providing cost-effective predictions for engineering flows with moderate computational resources. For higher fidelity, (LES) adapts FVM on unstructured grids to resolve large-scale eddies while modeling subgrid-scale effects, enabling detailed capture of unsteady turbulent structures in complex geometries. Practical implementations of FVM in CFD are exemplified by simulations of flows and chambers, leveraging open-source and commercial codes. In analysis, FVM codes like and Fluent compute and by resolving boundary layers and wakes around geometries such as the NACA series, often incorporating RANS for turbulent conditions at Reynolds numbers up to 10^6. For chambers, FVM simulates reacting flows in or injectors, modeling transport and release with finite-rate on polyhedral meshes to predict flame stabilization and efficiency. Multiphase flows are effectively handled within the FVM framework using the volume-of-fluid (VOF) method, which tracks interfaces by advecting a scalar across control volumes while solving a single set of Navier-Stokes equations for all phases. VOF ensures sharp interface representation and mass conservation, making it suitable for free-surface or droplet simulations in CFD, such as wave breaking or . Code verification in FVM-CFD relies on standard benchmarks like the lid-driven cavity, which tests incompressible solvers for recirculation at Reynolds numbers from 100 to 1000, and the backward-facing step, validating separation and reattachment in turbulent flows with expansion ratios of 1:2.

Other engineering and scientific domains

The finite volume method (FVM) has been extensively applied to heat and mass transfer problems, particularly in conduction-convection scenarios within porous media, where it ensures conservation of energy and species across control volumes. In simulations of heat transfer in porous media, FVM discretizes the governing equations for coupled conduction and convection, enabling accurate modeling of temperature and concentration fields in applications like fuel cells and thermochemical storage. For instance, a finite volume approach has been used to model heat and mass transfer in metal hydride-based hydrogen storage systems, capturing the interplay between thermal diffusion and reaction kinetics with high fidelity. Additionally, in radiative heat transfer, FVM solves the radiative transfer equation by integrating intensities over control volumes, providing robust solutions for participating media such as gases or semitransparent solids, as demonstrated in early seminal work by Raithby and Chui that established its efficiency for non-gray radiation models. In , FVM addresses challenges in nearly incompressible elasticity and poroelasticity through formulations that combine its conservation properties with finite element methods (FEM) for enhanced stability. For poroelastic media, mixed FVM-FEM simulate coupled deformation and fluid flow in heterogeneous materials, such as fractured reservoirs, by discretizing the Biot equations over control volumes while enforcing at interfaces. This approach mitigates volumetric locking in nearly incompressible regimes, as seen in multiscale FVM implementations that resolve elastic deformation under poroelastic loading with adaptive refinement. A FVM variant for further ensures across cell faces, making it suitable for static and dynamic problems in deformable solids without requiring structured grids. For electromagnetics, FVM discretizes to model wave propagation, leveraging its flux-conserving nature to maintain physical accuracy in time-domain simulations. The Yee scheme, a structured-grid FVM variant on staggered arrangements, approximates electromagnetic fields by integrating curls over dual control volumes, enabling stable propagation of waves in isotropic media and serving as the foundation for the finite-difference time-domain (FDTD) method. Extensions to unstructured grids generalize this approach, allowing Yee-like schemes to handle complex geometries for applications like design and photonic devices. Unstructured FVM formulations for further improve flexibility, reducing numerical dispersion in transient electromagnetic problems compared to traditional methods. In astrophysics, FVM facilitates radiative transfer modeling by solving the transfer equation on adaptive grids, capturing photon transport in stellar atmospheres and supernovae remnants with conserved energy. An improved FVM for radiative hydrodynamics integrates radiation source terms directly into hydrodynamic control volumes, enhancing accuracy for multigroup opacity treatments in high-energy astrophysical flows. For geophysics, FVM simulates seismic wave propagation on adaptive unstructured meshes, accommodating heterogeneous earth models for earthquake forecasting and exploration seismology. Godunov-type FVM schemes, in particular, provide monotone solutions for elastic waves, preserving positivity and reducing oscillations in velocity and stress fields across fault zones. Spectral FVM variants further optimize dispersion control on adaptive grids, enabling efficient modeling of wave scattering in layered media. Emerging applications of FVM include modeling, where it enforces conservation laws for atmospheric dynamics on cubed-sphere grids to simulate global circulation patterns. The GFDL finite-volume dynamical core, for example, integrates momentum and tracer with low numerical , supporting high-resolution forecasts of and variability. In simulation for , FVM resolves coupled transport equations for lithium-ion cells, tracking ion and electrochemical reactions across porous electrodes. Fully coupled FVM models predict cell performance during charge-discharge cycles, incorporating Butler-Volmer kinetics and porous media effects to optimize design parameters like capacity fade.

References

  1. [1]
    [PDF] Finite Volume Methods - UCI Mathematics
    The finite volume method (FVM) is a discretization technique for partial differential equations, especially those that arise from physical conservation laws. ...
  2. [2]
    2.5 Introduction to Finite Volume Methods | Unit 2 - DSpace@MIT
    The basis of the finite volume method is the integral convervation law. The essential idea is to divide the domain into many control volumes.
  3. [3]
    12 General Computational Fluid Dynamics
    During the 1970s upwinded finite difference, finite volume, and finite element methods were developed, leading to the creation of several industrial codes.
  4. [4]
    Finite volume methods for two-dimensional incompressible flows ...
    An example of a finite volume method ... S.V. Patankar. Numerical Heat Transfer and Fluid Flow. Hemisphere Publishing Corporation, McGraw-Hill, New York (1980).
  5. [5]
    Finite volume methods - ScienceDirect.com
    This chapter focuses on finite volume methods. The finite volume method is a discretization method that is well suited for the numerical simulation of ...
  6. [6]
    Finite Volume Methods for Hyperbolic Problems
    Similar to the book of R. J. LeVeque titled Numerical Methods for Conservation Laws, this manuscript will certainly become a part of the standard literature ...
  7. [7]
    [PDF] Comparison of node-centered and cell-centered unstructured finite ...
    Cell-centered and node-centered approaches have been compared for unstructured finite-volume discretiza- tion of inviscid fluxes.
  8. [8]
    (PDF) Origins and Development of the Finite Volume CFD Method at ...
    May 27, 2018 · The FVM started as an extension of the FDM from the integral formulation of the conservation equations. This work was led mainly by the ...
  9. [9]
    Numerical Heat Transfer and Fluid Flow | Suhas Patankar
    Oct 8, 2018 · (1980). Numerical Heat Transfer and Fluid Flow (1st ed.). CRC Press ... Using simple algebra and elementary calculus, the author develops ...Missing: finite volume
  10. [10]
    [PDF] Third-Order Accurate, Second-Derivative-Free Finite-Volume Method
    In this paper, we report progress in the development of a third-order accurate, second- derivative-free, shock-capturing finite-volume solver for ...Missing: advancements | Show results with:advancements
  11. [11]
    [PDF] Conservation Laws and Finite Volume Methods Notes
    Jan 3, 2011 · Finite volume Methods. • Approximate cell averages: Qn i ≈. 1. ∆x Z xi+1/2 x i−1/2 q(x, tn) dx. • Integral form of conservation law,. ∂. ∂t Z.
  12. [12]
    [PDF] Governing Equations of Fluid Dynamics
    Equations (2.42a–c) are the Navier-Stokes equations in conservation form. In the late seventeenth century Isaac Newton stated that shear stress in a fluid ...
  13. [13]
    On the freestream preservation of finite volume method in curvilinear ...
    Apr 28, 2016 · Finite volume method in curvilinear coordinates​​ Consider the three-dimensional hyperbolic systems in conservative form.
  14. [14]
    Galilean invariance and the conservative difference schemes for ...
    Galilean invariance for general conservative finite difference schemes is presented in this article. Two theorems have been obtained for first- and ...Missing: frame- | Show results with:frame-
  15. [15]
    E pur si muove: Galilean-invariant cosmological hydrodynamical ...
    The mesh is used to solve the hyperbolic conservation laws of ideal hydrodynamics with a finite-volume approach, based on a second-order unsplit Godunov scheme ...
  16. [16]
    Finite Volume Method — FiPy 3.4.5 documentation - NIST Pages
    Jun 26, 2024 · In the FVM the variables of interest are averaged over control volumes (CVs). The CVs are either defined by the cells or are centered on the vertices.
  17. [17]
    [PDF] Geometry Modeling, Flow Domain Modeling, and Grid Generation
    A structured grid can only contain finite-volume cells with a hexahedral ... Each approach has its advantages and disadvantages, but if things are done ...
  18. [18]
    [PDF] Finite Volume Method: A Crash introduction - Wolf Dynamics
    The FVM divides a domain into control volumes, seeks solutions inside, and uses the Gauss theorem to convert volume integrals to surface integrals.
  19. [19]
    An improved orthogonal grid generation method for solving flows ...
    Nov 1, 2014 · They observed that severe mesh skewness restricts allowable cell Peclet number, time-step size and adversely affects the phase and amplitude ...
  20. [20]
    An adaptive gradient correction method based on mesh skewness ...
    Jan 3, 2025 · In fact, mesh skewness is closely related to other mesh quality metrics, such as element orthogonality and aspect ratio. Higher skewness ...<|separator|>
  21. [21]
    [PDF] Implementation of Implicit Adaptive Mesh Refinement in an ...
    This paper explores the implementation of adaptive mesh refinement in an unstructured, finite-volume solver. Unsteady and steady problems are considered.
  22. [22]
    [PDF] Spectral (Finite) Volume Method for Conservation Laws on ...
    The original. Godunov scheme employed a piece-wise constant data reconstruction, and the exact Riemann solver, and the resultant scheme was only first-order.
  23. [23]
    Approximate Riemann solvers, parameter vectors, and difference ...
    References · 1. P.L. Roe. Proceedings, Seventh Int. Conf. · 2. O.A. Oleinik. Trans. · 3. P.D. Lax. Hyperbolic Systems of Conservation Laws and the Mathematical ...Missing: citation | Show results with:citation
  24. [24]
    (PDF) On the Discretization of the Diffusion Term in Finite-Volume ...
    The aim of this article is twofold: first, to present and analyze various practices for the finite-volume discretization of the diffusion term in continuum ...
  25. [25]
    [PDF] Chapter 13 The Transient Term - AUB
    This textbook explores both the theoretical foundation of the Finite Volume Method. (FVM) and its applications in Computational Fluid Dynamics (CFD).Missing: Nicolson | Show results with:Nicolson
  26. [26]
    None
    The document does not discuss low-storage Runge-Kutta schemes for time integration in CFD, their order of accuracy, or stability. It focuses on inelastic finite element analyses for structural components under cyclic thermal/mechanical loadings, detailing:
  27. [27]
    [PDF] Conservation Laws and Finite Volume Methods
    Jan 19, 2011 · The CFL Condition: A numerical method can be convergent only if its numerical domain of dependence contains the true domain of dependence of the ...
  28. [28]
    [PDF] Stability of a Crank-Nicolson Pressure Correction Scheme ... - HAL
    May 3, 2013 · The developed numerical scheme, based on a pressure correction technique, uses a Crank-. Nicolson time discretization and a staggered space ...
  29. [29]
    [PDF] Application of Dual Time Stepping to Fully Implicit Runge Kutta ...
    This paper presents the formulation of a dual time stepping procedure to solve the equations of fully implicit. Runge-Kutta schemes.
  30. [30]
    [PDF] The Finite Volume Method in Computational Fluid Dynamics
    The first source was the bi-yearly computational fluid dynamics (CFD) course, which has been offered over the last 15 years at the American University of Beirut.
  31. [31]
    [PDF] Implementation of boundary conditions in the finite-volume pressure ...
    May 2, 2016 · For example on a physical “Wall”, the boundary condition for the momentum equation could be a slip or a no-slip condition, while for the energy ...Missing: adjustment | Show results with:adjustment
  32. [32]
    5. Boundary conditions — Finite Volume method
    Finite Volume method : from 1D to 2D · Docs »; 5. Boundary conditions; View page ... Figure 5.1 West Boundary condution with a ghost cell (I=0). Let's ...
  33. [33]
    [PDF] Finite volume method on unstructured grids - Projects at Harvard
    Apr 1, 2013 · Dashed line represents the integration path of the Green-Gauss gradient evaluation (control volume gtt). ... In the finite volume method ...
  34. [34]
    [PDF] A Far-Field Non-re ecting Boundary Condition for Two-Dimensional ...
    Aug 24, 1995 · Far- eld boundary conditions for external ow problems have been devel- oped based upon long-wave perturbations of linearized ow equations ...
  35. [35]
    Finite Volume Methods for Hyperbolic Problems
    In this chapter we begin to study finite volume methods for the solution of conservation laws and hyperbolic systems. The fundamental concepts will be ...<|control11|><|separator|>
  36. [36]
    [PDF] Finite volume met hods: foundation and analysis
    Finite volume methods are discretization schemes used in fluid mechanics, meteorology, and other areas, known for numerical robustness and local conservation.Missing: core seminal papers
  37. [37]
    Finite Volume Method - an overview | ScienceDirect Topics
    The Finite Volume Method (FVM) is a numerical technique widely used for solving partial differential equations, particularly those arising from conservation ...
  38. [38]
    Error estimate for the upwind finite volume method for the nonlinear ...
    In this paper we estimate the error of upwind first order finite volume schemes applied to scalar conservation laws. As a first step, we consider standard ...
  39. [39]
    A Second-Order Maximum Principle Preserving Finite Volume ...
    The method is formally second-order accurate by means of a piecewise linear reconstruction within each cell and at mesh vertices. An algorithm is provided to ...
  40. [40]
    [PDF] implicit schemes and parallel computing in unstructured grid cfd
    This report outlines implicit schemes for steady state solutions, explicit/implicit schemes for unsteady flows, and parallelizing finite volume schemes on  ...
  41. [41]
    Multigrid Strategies for Viscous Flow Solvers on Anisotropic ...
    Unstructured multigrid techniques for relieving the stiffness associated with high-Reynolds number viscous flow simulations on extremely stretched grids are ...
  42. [42]
    A priori grid quality estimation for high-order finite differencing
    Jun 15, 2016 · ... accuracy, for three common geometric features found in grids: abrupt changes in the grid metrics, skewness, and grid stretching. A strong ...Missing: skewed | Show results with:skewed
  43. [43]
    [PDF] Solution methods for the Incompressible Navier-Stokes Equations
    Finite Volume (FV) discretization. Discretization of space derivatives ... Discretize the equations in conservation (integral) form. Eventually this ...
  44. [44]
    [PDF] Chapter 16 Fluid Flow-Compressible Flows - AUB
    This textbook explores both the theoretical foundation of the Finite Volume Method. (FVM) and its applications in Computational Fluid Dynamics (CFD).
  45. [45]
    [PDF] Turbulence Modeling of Shock Boundary Layer Interaction in ...
    Dec 1, 2021 · The turbulent models include Reynolds-Averaged Navier-Stokes (RANS) with 4 different closures (k-ε, k-ω, Shear Stress Transport (SST), and ...
  46. [46]
    [PDF] Large Eddy Simulation Using an Unstructured Mesh Based Finite ...
    This paper concerns development of a large eddy simulation (LES) capability for a finite-volume solver which permits use of unstrcutured meshes.
  47. [47]
    Introduction to Advanced Computational Methods - Eagle Pubs
    ... history of computational fluid dynamics (CFD). It was one of the first flow equations that researchers solved numerically using finite-difference ...
  48. [48]
    [PDF] CFD Simulation of Explosions in Fired Combustion Chambers
    The code is based on the finite volume method, and it solves the Reynolds Averaged Navier-Stokes (RANS) transport equations, including mass, momentum ...
  49. [49]
    [PDF] On The Design, Implementation, and Use of a Volume-of-Fluid ...
    In this dissertation we describe the design, implementation, and use of a two-dimensional, second-order accurate volume-of-fluid interface tracking ...
  50. [50]
    [PDF] Numerical Simulation of the Navier-Stokes Equations using Finite ...
    Among them, finite volume method is a relatively popular one. It divides the domain into a set of control volume, every control volume is represented by a ...
  51. [51]
    "Turbulent Heat Transfer Over a Backward-Facing Step: Solution to ...
    This paper presents a numerical solution to a two-dimensional turbulent convection flow over a backward facing step. The work is done in response to a technical ...
  52. [52]
    Multiscale finite volume method for finite-volume-based simulation of ...
    We propose a multiscale finite volume method (MSFV) for simulation of coupled flow-deformation in heterogeneous porous media under elastic deformation.
  53. [53]
    [PDF] hybrid stress finite volume method for linear elasticity problems
    Abstract. A hybrid stress finite volume method is proposed for linear elasticity equations. In this new method, a finite volume formulation is used for the ...Missing: poroelasticity | Show results with:poroelasticity
  54. [54]
    A Finite-Volume Method for the Maxwell Equations in the Time Domain
    A finite-volume scheme for the Maxwell equations is proposed using collocated nonorthogonal curvilinear grid arrangements, with the advantage that all ...
  55. [55]
    (PDF) Yee's scheme for the integration of Maxwell's equation on ...
    PDF | The co-volume integration method, Yee's scheme, generalized to unstructured mesh is considered and compared with time domain finite element method.
  56. [56]
    [PDF] Improved Finite-Volume Method for Radiative Hydrodynamics ( )
    It is most efficient to solve the radiative transfer equation on the same grid as the CFD solution, and this work presents an algorithm with improved accuracy ...
  57. [57]
    Novel adaptive finite volume method on unstructured meshes for ...
    Jul 1, 2023 · A new adaptive finite volume method is proposed for the simulation of the wave problems in the time domain.
  58. [58]
    On Godunov-type finite volume methods for seismic wave propagation
    May 30, 2025 · In this study, we delved into two Finite Volume (FV) methods to assess their efficacy and competitiveness in seismic wave simulations.
  59. [59]
    A new spectral finite volume method for elastic wave modelling on ...
    Abstract. In this paper, we consider a new spectral finite volume method (FVM) for the elastic wave equations. Our new FVM is based on a piecewise constant.
  60. [60]
    [PDF] A Scientific Description of the GFDL Finite-Volume Cubed-Sphere ...
    Jun 14, 2021 · This document is a scientific description of the GFDL Finite-Volume Cubed-Sphere Dynamical Core, including its introduction, solver outline, ...
  61. [61]
    [PDF] Fully Coupled Simulation of Lithium Ion Battery Cell Performance
    This work develops and implements a fully-coupled finite volume methodology for the simulation of the electrochemical equations in a lithium-ion battery cell.