The Art and Science of Meshing Airfoil

Figure 1: Blocking strategy to capture physics around an airfoil.

2098 words/ 11 minute Read

CFD 101 starts with airfoils. This geometry is seen as the stepping stone in the aerospace/turbomachinery field, before diving deep into CFD. Irrespective of the gridding software and the gridding methodology adopted to mesh, everyone, is called to achieve this goal of meshing an airfoil first.

Despite the fact that meshing strategies for an airfoil have come a long way, newer, smarter, and more efficient strategies continue to evolve, to capture the subtle physics in the most accurate and optimal way.

This article focuses on the traditional gridding strategies along with improved blocking techniques around airfoils for optimal CFD results.

H-type pattern for a bi-convex airfoil
Figure 2: H-type pattern for a bi-convex airfoil

H-type pattern for a curved leading edge airfoil, 3b - H-type pattern with a leading-edge split
Figure 3a:
H-type pattern for a curved leading edge airfoil, 3b – H-type pattern with a leading-edge split

H-type pattern with boundary layer clustering around an airfoil
Figure 4: H-type pattern with boundary layer clustering around an airfoil.

The Conventional Ways of Airfoil Meshing!

H-Type

One of the classical gridding approaches is the H-type pattern. It has finely clustered cells at the leading and trailing edges as shown in figure 2. This pattern is simple to construct and holds good for a biconvex airfoil with sharp leading and trailing edges. However, when applied to an airfoil with a curved leading edge, the H-pattern creates a singular point (figure 3a). This can partly be alleviated by splitting the singularity into two weaker singularities as shown in figure 3b. But still, it falls short of capturing the leading edge curvature accurately and also propagates the boundary layer in the transverse direction (figure 4).

Alongside the wasted use of point clusters and high aspect ratio cells, the high clustered cells both parallel and perpendicular in regions where the flow accelerates can result in significant time step reductions due to CFL conditions, leading to slowing of solver convergence.

C-type pattern for a blunt leading edge airfoil with boundary layer clustering
Figure 5: C-type pattern for a blunt leading edge airfoil with boundary layer clustering.

C-Type

An improved variant of H-type topology is the C-type pattern, which captures the leading edge curvature without any singularities. Though the C-type pattern avoids the propagation of boundary layer fineness upstream, it fails to do so downstream of the airfoil trailing edge. In a way, this downstream fineness proves beneficial as it helps to capture the shear layer for solver runs at low angles of attack.

In practice, however, most applications of the C-type pattern, have not taken advantage of the wonderful alignment of the grid lines along with the shear layer. To take full advantage of this pattern, the CFD analysis would have to continually shift the grid curves to dynamically align with the shear layer and this would result in greater conformity to the flow physics. While not doing this alignment results in algorithmic simplicity, it does result in the wastage of grid points downstream and falls short of being an efficient grid type.
O-type pattern for a blunt leading edge airfoil with boundary layer clustering

Figure 6: O-type pattern for a blunt leading edge airfoil with boundary layer clustering.

O-Type

The next classical blocking strategy is the O-type, which almost overcomes the disadvantages of the H-type and C-type grids. In O-type the entire grid sheet is wrapped up around the airfoil without any propagation of the boundary-layer fineness into the field. This helps in getting an optimal cell count, eliminating the redundant propagation of cells as seen in the H-type and C-type patterns.

However, this pattern has its inherent limitations. It creates highly skewed cells for zero thickness trailing edge airfoils. The grid cell on the upper airfoil surface directly connects to the cell on the lower airfoil surface across the horizontal grid line emanating from the trailing edge. For a single cell-to-cell step, it nearly creates a 360-degree turn in one step. Thus, each cell will have an angle of slightly less than 180 degrees which is not exactly of good quality.

The level of skewness created in such cases is acceptable for Euler grids but is extremely high in viscous grids. Singular points like the trailing edge are critical for accurate CFD computations. High skewness will affect the solver’s robustness and also the quality of the solutions. For this reason, many prefer to go with the C-type grid and live with the excessive cells in the wake region.


Cartesian grid. b - Unstructured gridFigure 7a: Cartesian grid. Image source – hanleyinnovations, 7b – Unstructured grid

Unstructured and Cartesian

In the conventional unstructured or cartesian grids, the approach is the same as O-type in the near vicinity of the wall. A few rows of cells are wrapped around the airfoil with the rest of the domain filled with unit aspect ratio tria’s or quad cells. Similar to the O-topology approach, the cells in the viscous padding at the trailing edge singularity point are skewed, deviating from the ideal requirement of high-quality cells.

In some variants of boundary layer creation, such as seen in figure 8, the trailing edge singularity point can be captured without any skewness or loss of orthogonality. These are ideal grids with optimal cell count. But when the need to capture the wake arises, the Cartesian or the unstructured approach results in a humongous increase in cell count, because of the inherent limitation of generating unit aspect ratio cells.

Trailing edge singularity capturing with unit aspect ratio cellsFigure 8: Trailing edge singularity capturing with unit aspect ratio cells.

A Quintessential Mesh

Having gone through the pros and cons of the traditional meshing strategies, you might be wondering what does an ideal grid for an airfoil look like? Let’s have a look at the gridding requirements for an airfoil and see what the geometric features and the physics demand.

A single element airfoil needs a sufficient number of points on the body, say typically in the range of 300-400 points for RANS simulation to capture the geometry. A finite thickness trailing edge demands at least 20-30 points to maintain adequate grid density around the singularity region. The point placement is of the order of 0.1 percent of chord length at the leading and trailing edges, with gradual sparsening towards the mid chord region.

At the least, 20-40 layers of stretched elements are stacked orthogonally to the wall to capture the boundary layer, with a first cell height placement estimated for a Y-plus of ~1.0 based on the given flow Reynolds Number. Usage of a smaller stretching ratio, say lesser than 1.2 is highly recommended in building the viscous padding. More the orthogonality in these layers, the better the CFD results.

In unstructured and Cartesian gridding approaches, one needs to pay additional attention to the cell height ratio of the last layer of the boundary layer padding and the elements thereafter into the domain. The size variations need to be as small as possible. This is a critical aspect influencing the solver’s robustness and solution accuracy. However, this is not an issue in structured grids as the underlying gridding philosophy ensures smooth transitions.

Next, the grid outside of the viscous padding should grow gradually till they reach the farfield. Keeping the farfield as far away as possible is strongly recommended, to ensure that they don’t influence the solution in the near vicinity of the airfoil. For a single-element airfoil, the farfield needs to be at least 30 chords away in all directions. NASA’s website for Turbulence Model Verification recommends keeping the farfield as far away as 500 chords. This is too far than actually needed, but they recommend it for sequential grids generated for grid convergence study. For day-to-day industrial production runs, this is not a must, as the solution one obtains from a 30-100 chord farfield distance is of sufficient quality and accuracy.

Limitations

The above specifications give an idea of a basic mesh for an airfoil, fulfilling the need to accurately represent the geometry and capture the gross physics. This basic grid gives reasonably good results and in most instances, this is the grid one settles down for routine production runs.

But, a lot more could be done for capturing the rich flow physics around an airfoil. A C-type, H-type grid can capture the downstream wake for low alphas with smooth streamline flow around the airfoil body with no flow separation. Under, flow conditions with rich physics like large flow separations and discontinuous flow phenomenons like shock or expansion fans, this basic grid falls short in accurately capturing them.

The “Holy Grail” of  Meshing?

Adaptivity looks to be the ideal choice to accurately capture flow phenomenons. By definition, adaptivity refers to local refinement of the flow domain respecting the physics. Refinement can either be solution-based or selective local refinement based on the Engineer’s intuitive feel of the flow physics.

Both adaptive strategies have their pros and cons. Solution-based adaptivity on the positive side is excellent in the selective refinement of regions rich in flow physics, ignoring regions with less activity. On the downside, this demands multiple intermediate CFD solver runs before settling down to the final grid. Also, adaptation runs the risk of placing hanging nodes on the body, which hampers the robustness of many CFD solvers.

On the other hand, the user-defined local refinement using various field refinement options in the grid-generator helps to refine the important regions with one-to-one cell connectivity. Though the approach is limited by the intuitive feel of the flow physics by the engineer, in many instances it end-up with a few additional cells. Nevertheless, this is still a widely used approach with the inherent benefit of one grid catering to the needs of multi-angles of attack simulations, without the intermittent solver runs.

Topology for shock capturing on the upper surface of the airfoilFigure 9: Topology for shock capturing on the upper surface of the airfoil.

New Efficacy of Multi-block Meshing

Major flow phenomenons like the wake and shocks can be effectively captured by smart topology building without the limitations posed by traditional-blocking strategies like the C- type and O-type topologies. Figure 9 shows the grid block layout for shock capturing with a known shock position. As it can be observed, the blocks are placed right at the discontinuity, capturing the shock with a dense cloud of grid points and gradually smoothing out as we move away from the location. The grid density in the central cocoon of blocks can be varied to the user’s desired level of resolution, without affecting the neighboring block grid size.

O-Type grid for an Airfoil with Blunt Leading and Trailing Edges with trailing edge and upper surface resolution plus shear layer resolution with a sleeve surface to contain the anticipated shear layerFigure 10: O-Type grid for an Airfoil with Blunt Leading and Trailing Edges with trailing edge and upper surface resolution plus shear layer resolution with a sleeve surface to contain the anticipated shear layer.

Another example is shown in figure 10, where the wake from the airfoil is accurately captured, by using a loop topology. Stretched high aspect ratio cells are used, which judiciously capture the wake with optimal cell count. By extending the width of the loop, a larger region engulfing the upper part of the airfoil and the region downstream can be captured, extending the benefit of wake capturing and flow separation even at high angles of attack. The pre-eminence of this blocking strategy is that it avoids the propagation of mesh fineness further downstream, aiding in generating an optimal grid. Unlike the C- type grid, the fineness of the boundary layer does not propagate downstream, the aspect ratio of the cells inside the loop is under the user’s control, it is flexible to meet users varying requirements and hence less prone to code-blow outs.

A more effective technique of wake and flow separation capturing is the topology building concept of “nesting” as shown in figure 11. This, not only has all the benefits of the above-discussed looping strategy but also has the added advantage of rapid cell count reduction in the direction normal to the wake. This serves as a very powerful tool to creatively capture any mid-field flow physics without the humongous increase in cell count.

O-Type topology for an Airfoil with Blunt Leading and Trailing Edges with a sleeve surface to contain the anticipated shear layer plus nested cross-stitching for local downstream cross-sectional enrichmentFigure 11: O-Type topology for an Airfoil with Blunt Leading and Trailing Edges with a sleeve surface to contain the anticipated shear layer plus nested cross-stitching for local downstream cross-sectional enrichment.

Let Bygones be bygones

In bygone days, if CFD was able to capture the gross flow physics and make a modest prediction of the flow parameters, it was perceived as an accomplishment. CFD now is no more seen as an aid to Experiments for data cross-verifications but is widely accepted as a precursor and a reliable design tool.

The innovative strategies discussed here are not out of context. They are very effective and they enhance the CFD results, making them more reliable and accurate. Recent advances in computer technology have paved the way for transitioning from RANS to higher-order CFD methods, LES, and even DNS. And they all demand low aspect ratio, high-quality quad/hex meshes. These unique blocking strategies help to cater to their needs, without the hidden cost of cell count shoot-up, loss of flow-alignment and orthogonality, poor cell quality as seen by conventional gridding techniques.

Further Reading

  1. Is Grid Generation an Art or Science?
  2. The Art and Science of Meshing Turbine Blades
  3. Grid Convergence Study ! – Is it Necessary?
  4. Do Meshes still play a critical role in CFD ?
  5. Nesting your way to mesh Multi-Scale CFD Simulation!

 

Twitter
Visit Us
Follow Me
YouTube
LinkedIn
Share

Subscribe To GridPro Blog

By subscribing, you'll receive every new post in your inbox. Awesome!

4 Comments on “The Art and Science of Meshing Airfoil”

  1. Great article, very helpful.
    I wonder if you have some references (articles or literature) on this subject (meshing airfoils) that you judge as relevant. I am currently working on this and being able to reference it would be great.

    1. Hi Lucas,
      We appreciate you reading through the article!
      It was written based on our expertise in meshing airfoils using GridPro. So we do not have any paper/material which we can recommend for this currently.

  2. Thank you for explaining the limitations and advantages of the different mesh grids, especially transonic cases.

    I am really awe-felt by looking at these amazing mesh grids! I think I can’t do this in 3D, let alone 2D. WOW!

  3. Hi Heich,
    Happy to know the article has helped you in having a better understanding of grid generation.

    Structured multi-blocking in Gridpro is relatively easy both in 2D as well as in 3D. In 2D it is a lot easier. In 3D also with experience, we can build complex nesting structures.

    You can just download Gridpro and give it a try. With the help of the tutorials in the Gridpro Youtube channel you can quickly get a hang of building these structured grids.

    with warm regards,
    Ravindra

Leave a Reply

Your email address will not be published. Required fields are marked *