0% found this document useful (0 votes)
41 views19 pages

Airfoil Mesh

Uploaded by

zain
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
41 views19 pages

Airfoil Mesh

Uploaded by

zain
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 19

Fluid Mechanics -II

4th Semester
NACA 64A-010 Airfoil
Project Report

Submitted to: Dr Zaib Ali

Name CMS

Muhammad Azhaf Zeeshan 423683


Analysis Of NACA 64A-010 Airfoil Through CFD
Muhammad Azhaf Zeeshan1,
¹SMME, NUST, Pakistan
1
[email protected]

Abstract— The following report below gives a CFD analysis of the NACA 64A-010 airfoil to investigate its
aerodynamic characteristics analytically. The NACA 64A-010 is a non-symmetrical airfoil with a 10%
thickness-to-chord ratio that is commonly utilised for its low-speed performance and in huge wind turbine
blades. During the CFD simulation on Ansys Workbench R2 2023 an airfoil was created, on which a mesh was
built, the governing equations were solved with a turbulence model, and finally the solution was visualized.
Pressure distribution over the airfoil surface was used to calculate lift and drag forces. By analysing the flow
velocity and contour maps, an in-depth understanding of the airflow around the airport could be used to
pinpoint crack locations. Lift and drag coefficients were calculated to evaluate the performance of the NACA
64A0 To ensure accuracy the findings were compared with previously collected experimental data or
experimental solutions from various sources collected from the internet. Although it can it has been difficult
to obtain an exact match, mesh and the simulation parameters were changed several times in order to obtain
the most realistic results.

I. INTRODUCTION
An airfoil is a form or cross-section of an item, such as a wing, blade, or turbine, that generates aerodynamic
force when moving through a fluid (e.g. air or water). An airfoil shape in an aeroplane generates lift by creating a
pressure differential between the top and lower surfaces when air flows over them. Airfoils are utilised in a variety
of applications, including aeroplanes, wind turbines, hydrofoils, and propellers. [1]

Figure 1:General Airfoil with Labelling

The shape of an airfoil performs an essential element in its flight. The leading edge is first and for most factor,
at the same time the trailing edge marks the back of the airfoil. These surfaces converge at both ends. The distance
among them is the airfoil's chord, that may vary along a wing's span resulting in different chord lengths. A key
layout function is camber, that's the curvature of the upper floor relative to the flatter lower one. This asymmetry
generated through the shape is what allows generate carry by means of deflecting airflow over the top for a longer
direction, growing a stress distinction. The camber line, imagined because the centreline between the higher and
decrease surfaces, enables define this curvature.

I-A: NACA-64A-010 AIRFOIL


The symmetrical NACA 64A010 airfoil, produced by the National Advisory Committee for Aeronautics,
changed into an essential milestone in the area of aerodynamic studies. It is symmetrical because the higher and
lower surfaces of the airfoil are identical. Because of its accurately documented homes and simplicity of
geometry, NACA 64A010 can serve as benchmark and device for scientists in the research area. In research, it's
far used for validating Computational Fluid Dynamics (CFD) models and wind tunnel experimental data. By
analysing its behaviour, scientists can affirm their computational models and have a better knowledge of shape -
mainly camber and thickness - effect on lift and drag forces for an airfoil. But the symmetrical airfoil of NACA
64A10 is limited in its ability to provide lift, making it not an airfoil that can be used for real flight. The simple
design of the 64A010 airfoil means that many people learning about airfoil layout and aerodynamics can build
the necessary knowledge. The symmetrical geometry simplifies the complexities of the airfoil design, providing
a clear know-how of the airfoil performance and camber and thickness impact. But the statistics learned from
NACA 64A010 are vital for the improvement of a plethora of airfoils utilized in airplanes, wind generators, and
other gear. [2]

Figure 2: NACA-64A-010 Airfoil

II: METHODOLOGY
For the purpose of analysing NACA 64A-010 Airfoil, to solve the RANS equations, the ANSYS Fluent
Student Version 2023 was employed. The Reynolds Averaged Navier Stokes equations (RANS), which are
derived from the fundamental principles of conservation of mass, momentum, and energy, were solved.
Ansys Student 2023 is a free program offered by Ansys that provides students with a powerful toolkit for
engineering simulations. Focused on Ansys Workbench, a central platform for various simulation tools, it offers
student-friendly versions of popular software like Ansys Mechanical for structural analysis and Ansys CFD for
fluid dynamics.
A. Importing of Coordinates
Ansys Fluent uses the Finite Volume Method a unique numerical technique used to solve the aforementioned
equations efficiently. The initial phase of this project started with the creation of a model of the NACA 64A-010
airfoil geometry within ANSYS DesignModeler software. This process began with acquiring the coordinate data
for the said airfoil from the National Advisory Committee for Aeronautics (NACA) website. The data was then
successfully imported into a Microsoft Excel spreadsheet, where the coordinates for the upper and lower airfoil
surfaces were separated. This separation ensured that an equal number of points were utilized for each surface,
thereby promoting an accurate representation of the airfoil's geometry. Following this meticulous organization,
the data was then imported into the ANSYS DesignModeler environment. Within DesignModeler, separate two-
dimensional (2D) curves were constructed to represent the upper and lower surfaces of the airfoil. The final step
in this initial phase involved the connection of these 2D curves, ultimately resulting in the generation of a complete
digital model of the NACA 64A-010 airfoil.

B. Defining Computational Domain


After the creation of the airfoil geometry, the following step included defining an airflow domain, which for
the sake of simplicity was taken to be a c shaped domain or a closed contour where the simulation would take
place. This domain was drawn around the airfoil and ensured the closing of the airfoil throughout the domain,
allowing it to experience the desired and controlled flow conditions for accurate analysis. The domain was then
divided into two parts and the inlet and outlet selection were then made in order to define the entrance and exit of
velocity in the said domain. For the following airfoil, the chord length was taken as 1m (for the purpose of
accordance with data sheet). The chord length taken was 1 min length, A line was sketched, at the trailing edge.
The sole purpose of sketching that line was to enhance the meshing to be carried out in the next step.
While the semi-circular region radius was 3 m, and the wall length was 5 m. Since the flow field was considered
a fluid for analysis, surface generation was a crucial step which was rendered by creating a closed surface using
Sketch command. Since the airfoil was to be analysed for the external properties, it meant that the airfoil was to
be subtracted from the generated surface, which was done so using the subtract Boolean command.

C. Meshing
Once the computational domain is defined, the next important step that followed was the mesh definition.
Meshing involves discretising or breaking down the domain into a network of small cells. These cell lay down the
foundation for CFD.The finer the cell network especially around the airfoil, the better and more accurate the
results gathered.. The generated cells/blocks exist both in 2D and 3D, specifically triangular and quadrilateral in
the 2D case and tetrahedral, hexahedral in 3D case. Considering broader aspect for optimization in case of NACA-
64A-010, fine meshing was considered with the triangular meshing type chosen. Keeping in view the fine meshing,
the entire contour was split into two regions for the sake of better meshing results.Edge sizing and face sizing was
done with their respective biasing factor in order to obtain a finer mesh that is more well defined around the
airfoil,especiallythe leading and trailing edged itself all for the sole purpose of obtaining more accurate mesh and
results

Figure 3: NACA-64A-010 Airfoil Triangular Mesh

III: SETTING UP SOLUTION


Once meshing has been completed successfully, it is required to set up an acceptable solution configuration.
Any analysis is conducted using an algorithm that specifies the required environment and conditions. The
pressure-based technique was created primarily for low-speed incompressible flows, whereas the density-based
approach was meant to be suitable for high-speed compressible flows. Analysis from a larger viewpoint
indicates that the methods have been expanded and redefine to solve and operate for a wide variety of flow
conditions beyond their standard or original implications, and the velocity field is obtained from the momentum
equations.

Figure 4: CFD Flowchart for pressure-based approach.


A. Boundary Conditions
1) Inlet Velocity
Inlet velocity is a critical aspect of boundary condition that defines the characteristics of airflow that
enters the domain. In the case of NACA-64A-010 the inlet velocity was varied in its x and y components for
different simulation runs. This allowed for analysing the airfoil behaviour under different flow directions and
magnitude. Along with that the velocity was also varied by choosing different Reynolds number for Angle of
Attacks which implored different flow regimes under the airfoil’s performance.

Figure 5: Selection of Inlet Velocity to change Re and AoA

2) Outlet Pressure
A pressure outlet boundary condition was used at the exit of the computational domain. This condition
specifies the static pressure at the outlet, allowing the flow to freely exit the domain. The pressure outlet condition
remained constant throughout the simulations for different angles of attack and Reynolds numbers.

3) Wall Conditions
Wall boundary conditions were applied to define the behaviour of the flow at the solid boundaries,
including the airfoil surface and the outer perimeter of the computational domain. In most cases, a no-slip wall
condition is used for viscous flows, which essentially states that the fluid velocity at the wall is zero. This reflects
the friction between the fluid and the solid surface. The wall conditions remained the same for all simulation runs
with different angles of attack and Reynolds numbers.

B. Turbulence Model Selection


The k-ω SST turbulence model was chosen for this CFD analysis due to its capability of handling the
anticipated flow characteristics around the airfoil. Airfoils can, especially around high angle of attacks
experience turbulent flow with unpredictable fluctuations. The k-ω SST turbulence model serves as an
excellent choice because it can handle a wide range of flow conditions including complex pressure changes
and boundary layer separation. This model offers a good balance between accuracy and computational cost,
making it ideal for complex geometries like airfoils where more advanced models can be too demanding. [1]
C. Adjusting External Parameters

Table 1:Adjusted Parameters and Values


Parameters Values
Analysis Type Steady State
Operating Temperature 23 ͦC
Operating Pressure 1x105 Pa
Air Density 1.225 kg/m3
Turbulence Model k-ω Model
Dynamic Viscosity (Air) 1.802x10-5 Pa.s

Residual monitoring was setup at 0.000001 and keeping in view the k-ω turbulence modelling, the following
parameters were outlined in accordance with the typical standards. Moreover, for smooth modelling, the number
of iterations was set initially to 1000.

III: RESULTS AND ANALYSIS


In conducting Computational Fluid Dynamics (CFD) analysis, a critical step involves establishing a suitable
solution environment to assess various solutions across different angles of attack and Reynolds Numbers.
Specifically, three distinct Reynolds Numbers were considered alongside their corresponding three angle of
attacks to showcase their behavior under given conditions. This analysis was conducted within the Fluid Flow
interface of Ansys Fluent.

Figure 6: Geometry Of NACA 64A-010

Figure 7 :NACA 64A -010 Mesh In Ansys Flow Fluent showing Inlet Outlet Selection
III-A: STANDARD FLOW CONDITIONS

For Standard Flow Conditions a suitable Re range was selected for different angle of attacks for which analysis
and calculation were run on Ansys Flow Fluent with a considerable amount of iterations done to get best possible
flow results,in case of our airfoil it was decided to be around 2-8*105 Re for best fit results under standard flow
conditions

Figure 8: Velocity Magnitude Bar for reference Figure 9: Static Magnitude Bar for reference

A) For Angle of Attack = 6°

Case Angle of Velocity Reynolds No CL CD


Attack
C1 6 120 8.158 * 105 0.6421 0.0157

Figure 10: Velocity Contour at 120m/s Figure 11: Pressure Contour at 120 m/s

1) Velocity Contour
As we can see there is higher velocity (orange) on the top of the airfoil which indicates that the flow is
accelerated due to reduced pressure. This is what generates lift. Conversely, the lower velocity (green) beneath
the leading edge airfoil indicates decelerated flow due to increased pressure. The orange at top and below the
body indicate high velocities while low velocity is seen at the trailing edge and at boundary layer due to
interaction of airfoil with the fluid(air)

2) Pressure Contour
As we can see, The yellow regions at both ends (leading and trailing edges) show areas of higher
pressure. This is typically where the air first contacts the airfoil (leading edge) and where it leaves the airfoil
(trailing edge). The blue region on the upper surface towards the leading edge indicates an area of lower
pressure. This is due to the Bernoulli’s principle, which states that an increase in the speed of a fluid occurs
simultaneously with a decrease in pressure.
Case Angle of Velocity Reynolds No CL CD
Attack
C2 6 30 2.039 * 105 0.6545 0.02269

Figure 12: Velocity Contour at 120m/s Figure 13: Pressure Contour at 120 m/s

1) Velocity Contour
As we can see there is higher velocity (orange) on the top of the airfoil which indicates that the flow is
accelerated due to reduced pressure. This is what generates lift. Conversely, the lower velocity (green) beneath
the leading edge airfoil indicates decelerated flow due to increased pressure. The orange at top and below the
body indicate high velocities while low velocity is seen at the trailing edge and at boundary layer due to
interaction of airfoil with the fluid(air) but are much less pronounced due to it being less velocity

2) Pressure Contour
As we can see, The yellow regions at both ends (leading and trailing edges) show areas of higher
pressure. This is typically where the air first contacts the airfoil (leading edge) and where it leaves the airfoil
(trailing edge). The blue region on the upper surface towards the leading edge indicates an area of lower
pressure. This is due to the Bernoulli’s principle, which states that an increase in the speed of a fluid occurs
simultaneously with a decrease in pressure, but are much less pronounced due to much less velocity hence
Reynolds number used

B) For Angle of Attack = 12°

Case Angle of Velocity Reynolds No CL CD


Attack
C3 12 10 6.797* 104 0.6942 0.08593

Figure 14: Velocity Contour at 10m/s Figure 15: Pressure Contour at 10 m/s
1) Velocity Contour
As we can see the contour is mostly green indicating not much resistance is caused and the contour is
generally undisturbed however the yellow spot on the top of leading edge indicates high velocity while the blue
outline alongside the border of airfoil shows lower velocities at boundary layer which is caused due to
obstruction by the airfoil.

2) Pressure Contour
As we can see,the red orange spot below the leading edge indicates high pressure and low velocity due
to first interaction of airfoil,the yellow spots around the leading and trailing edge indicates medium pressure
whereas the rest of it is pretty much unaffected this is mainly due to the low velocity chosen at this angle of
attack.

Case Angle of Velocity Reynolds No CL CD


Attack
C4 12 20 1.359* 105 0.6753 0.08757

Figure 16: Velocity Contour at 20m/s Figure 17: Pressure Contour at 20 m/s

1) Velocity Contour
Similar to the contour of 10 m the contour depicts streamwise velocity contours, with green signifying
minimal flow resistance and minimal distortion of the flow field. However, a localized region of elevated
velocity (yellow) exists near the leading edge on the upper surface. Conversely, a reduced velocity (blue) region
is observed adjacent to the lower leading edge and along the airfoil's boundary layer. This deceleration is
attributed to the presence of the airfoil, which introduces a viscous drag force that impedes the flow.

2) Pressure Contour
Mirroring the 10 m/s velocity contour, the pressure distribution exhibits minimal influence at most
locations, indicated by the prevalent neutral tones. However, a distinct region of elevated pressure (red-orange)
is evident below the leading edge. This corresponds to the initial interaction of the airflow with the airfoil,
resulting in stagnation and a subsequent decrease in velocity. Conversely, the yellow regions surrounding the
leading and trailing edges suggest moderate pressure levels. This can be attributed to the chosen angle of attack,
which, at this specific value, likely generates a less pronounced pressure differential across the airfoil compared
to higher angles.
Case Angle of Velocity Reynolds No CL CD
Attack
C5 12 30 2.039* 105 0.7748 0.06070

Figure 18: Velocity Contour at 30m/s Figure 19: Pressure Contour at 30 m/s

1) Velocity Contour
Similar to the 20 m/s case, the presented contour plot depicts streamwise velocity. Green hues signify
minimal flow resistance and negligible distortion of the flow field. However, a localized region of heightened
velocity (yellow) is evident near the leading edge on the upper surface while the size of localized regon is
significantly higher in this case due to higher velocity. Conversely, a region of diminished velocity (blue) is
observed adjacent to the lower leading edge and along the airfoil's boundary layer. This deceleration can be
attributed to the presence of the airfoil, which introduces a viscous drag force, thereby impeding the flow..

2) Pressure Contour
Similar to the observations at lower velocities, the pressure distribution at 30 m/s exhibits minimal
influence on most of the airfoil surface, as evidenced by the prevalence of neutral tones. However, a distinct
region of elevated pressure (red-orange) is observed below the leading edge. Notably, this region appears
slightly larger compared to the distributions at 10 m/s and 20 m/s. This intensification corresponds to the
intensified stagnation effect due to the higher flow velocity encountering the leading edge. Consequently, a
more pronounced decrease in velocity is expected in this region. Conversely, the yellow regions surrounding the
leading and trailing edges suggest moderate pressure levels.

C) For Angle of Attack = 25°

Case Angle of Velocity Reynolds No CL CD


Attack
C6 25 20 1.359* 105 1.1955 0.0967

Figure 20: Velocity Contour at 20m/s Figure 21: Pressure Contour at 20 m/s
1) Velocity Contour
The contour shows the presence of regions with elevated velocity (yellow and orange) situated above
and below the leading edge of the airfoil. Conversely, a region of reduced velocity (light and dark blue) is
observed on the upper surface progressing towards the trailing edge. This disparity in velocities suggests the
potential for flow separation in this region. The high angle of attack likely contributes to this phenomenon, as it
intensifies the adverse pressure gradient on the upper surface, potentially causing the boundary layer to detach
from the airfoil..

2) Pressure Contour
A prominent region of low pressure (blue) is evident on the upper surface near the trailing edge of the
airfoil. This signifies a zone of significant flow acceleration, consistent with Bernoulli's principle which dictates
an inverse relationship between fluid velocity and pressure. Conversely, high-pressure regions (yellow and red)
are observed at both extremities - on the lower surface near the leading edge and on both surfaces close to the
trailing edge. These areas correspond to flow deceleration, resulting in a pressure increase.

III-B: STALL FLOW CONDITIONS

At very high angles of attack (like 25 degrees) and low airspeeds, stall, a critical aerodynamic phenomenon can
occur at airfoils. The airflow struggles to follow the curved upper surface, causing the boundary layer to
separate prematurely typically around the mid-chord. This separation disrupts the pressure distribution, reducing
lift significantly. Low airspeeds further weaken the flow, worsening the stall. Maintaining sufficient airspeed is
crucial, especially during maneuver’s with high angles of attack.

For Angle of Attack = 25°

Case Angle of Velocity Reynolds No CL CD


Attack
C7 25 10 6.797* 104 1.1663 0.0797

Figure 22: Velocity Contour at 10m/s Figure 23: Pressure Contour at 10 m/s

1) Velocity Contour
There’s evidence of flow separation at the leading edge to the trailing edge, indicated by blue
coloration. This suggests turbulent or recirculating flow in that area due to the high angle of attack. Flow
separation occurs when the boundary layer travels far enough against an adverse pressure gradient that the speed
of the boundary layer relative to the object falls almost to zero.

The overall shape and colour gradient suggest that this is a simulation visualizing stall under the said
conditions. The flow separation and the resulting turbulent or recirculating flow can significantly affect the lift
and drag characteristics of the airfoil

2) Pressure Contour
Prominent high-pressure zones (orange and red) are observed on the lower leading edge and both
trailing edges of the airfoil. These regions correspond to flow deceleration and a subsequent pressure increase.
This creates a crucial pressure differential across the airfoil, with lower pressure acting on the upper surface and
higher pressure on the lower surface. this is a pressure contour indicating stall by colourization.
Case Angle of Velocity Reynolds No CL CD
Attack
C8 25 1 6.797* 103 1.1 0.0996

Figure 24: Velocity Contour at 1m/s Figure 25: Pressure Contour at 1 m/s

1) Velocity Contour
The prominent yellow-red coloration near the leading edge alongside the low velocity depicted by blue
coloration suggests flow separation and stall, a critical phenomenon at high angles of attack. This separation
arises when the boundary layer momentum can no longer overcome the pressure gradient pushing it against the
airfoil surface. Consequently, the airflow detaches, likely transitioning to turbulent or recirculating flow in this
region. This separation significantly disrupts the airfoil's performance, potentially leading to a stall. Stall is a
condition where the lift coefficient of the airfoil decreases due to airflow detachment from the wing's surface.
This phenomenon drastically reduces the wing's ability to generate lift, impacting an aircraft's control and
potentially leading to a loss of altitude.

2) Pressure Contour
A prominent yellow region near the leading edge signifies a high-pressure zone, indicative of airflow
deceleration due to the airfoil's shape. Conversely, a significant blue region on the upper leading edge reveals a
zone of low pressure. This pressure differential aligns with Bernoulli's principle, where faster-moving air (upper
surface) experiences lower pressure compared to the slower-moving air (lower surface) near the leading edge.
However, at high angles of attack, the pressure distribution can be significantly altered by flow separation. The
yellow coloration near the leading edge suggests flow separation, where the boundary layer detaches due to the
adverse pressure gradient. This separation can further enlarge the low-pressure region on the upper surface,
potentially triggering a stall – a critical condition where lift reduction occurs due to airflow detachment.

III-C: SHOCKWAVE CONDITIONS

A shockwave is a sort of propagating disturbance that travels faster than the local speed of sound in the medium.
Shockwave is an energy carrier that, like a regular wave, may travel in a medium but is distinguished by a rapid,
practically discontinuous change in pressure, temperature, and density of the medium. Unlike regular sound
waves, a shock wave's velocity fluctuates with its amplitude. A shock wave's speed is always greater than the
speed of sound in the fluid and decreases with wave amplitude. When the shock wave speed equals the standard
speed, the shock wave fades and transforms to a regular sound wave.
Figure 26: Shockwave generated by a typical supersonic aircraft

Case Angle of Attack Velocity Reynolds No


C9 6 500 1.46 Ma

Figure 27: Velocity Contour at 500m/s Figure 28: Pressure Contour at 500 m/s

1) Velocity Contour
The color gradient in the velocity contour profile serves as a visual representation of the supersonic
flow field (500 m/s) surrounding a NACA 64A-010 airfoil positioned at a 6-degree angle of attack. The
dominance of red and orange hues signifies regions of high velocity as the air accelerates past the airfoil.
Notably, the high speeds associated with supersonic flight lead to the formation of shock waves. These shock
waves arise due to the abrupt compression of air molecules exceeding the local speed of sound. As the airfoil
disrupts the surrounding air, localized pressure changes occur. When these pressure differences surpass the
sonic threshold, shock waves propagate outward, influencing various aspects of aerodynamics and propulsion.
Additionally, the phenomenon of sonic booms is directly linked to the formation of shock waves. Within the jet
plume emanating from the airfoil, a pattern of alternating light and dark bands emerges. These bands, also
known as shock diamonds or Mach disks, are a telltale signature of supersonic flow. Their presence signifies the
interaction of the exhaust flow with periodic shock waves. Each diamond represents a localized region where
the shock wave has induced a surge in pressure and temperature

2) Pressure Contour
This pressure contour visualizes the flow field surrounding a NACA 64A-010 airfoil at a supersonic
speed of 500 m/s and a 6-degree angle of attack. The color gradient depicts a map of pressure distribution, with
red signifying high pressure and blue indicating low pressure. A notable feature is the sharp transition in color
on the upper airfoil surface, indicative of a shockwave – a phenomenon arising from the object's supersonic
speed exceeding the local speed of sound. As the airfoil accelerates, it compresses air molecules rapidly, leading
to localized pressure variations. When these variations become significant enough to surpass the sonic threshold,
a shockwave forms and propagates outward, a signature of supersonic flight.

Following the shockwave, a fascinating pattern of alternating high- and low-pressure regions emerges. These
diamond-shaped features, known as shock diamonds or Mach disks, are another hallmark of supersonic flow,
often observed in jet engines or rocket exhaust plumes. Each diamond represents a localized area where the
shockwave has induced a surge in pressure and temperature. The periodic nature of these diamonds reflects the
continuous interaction between the supersonic flow and the surrounding atmosphere, resulting in a complex
interplay of pressure and temperature variations.

IV: DISCUSSION

IV-A: KEY FINDINGS

The analysis of the NACA 64A-010 airfoil revealed interesting relationships between velocity, angle of attack
(AoA), and the resulting pressure distribution and flow behavior.

A) Angle of Attack Effects:

1)Low AoA (6°):

At low AoA, the airfoil presents a minimal angle to the oncoming airflow. This allows the flow to remain
attached to the airfoil surface for most velocities (120 m/s, 30 m/s). Bernoulli's principle dictates that as air
accelerates over the curved upper surface (due to the AoA), it experiences lower pressure compared to the slower-
moving air beneath the airfoil. This pressure difference contributes to lift generation. The higher pressure observed
at the leading and trailing edges is due to stagnation and flow separation at these points, respectively.

2)High AoA (12°):

As the AoA increases (12°), the challenge for the airflow to follow the curved upper surface becomes more
significant. This is particularly true at lower velocities (10 m/s) where the airflow momentum is weaker. While
minimal effects might be observed initially, as velocity increases (20 m/s, 30 m/s), the pressure distribution reveals
a slight separation region near the leading edge on the upper surface. This separation occurs because the adverse
pressure gradient (pressure increasing in the flow direction) becomes too strong for the boundary layer to
overcome. This detachment disrupts the smooth airflow, reducing lift and potentially increasing drag.

3)Stall Conditions (25°):

At a very high AoA (25°), especially at low velocities (10 m/s, 1 m/s), the adverse pressure gradient becomes
too severe, leading to a complete flow separation across the airfoil. This phenomenon, known as stall, significantly
reduces lift. The pressure distribution for these stall cases shows a large low-pressure region on the upper surface,
indicating a significant portion of the airflow has detached from the airfoil.

B)Velocity Effects:

Higher velocities generally lead to more pronounced effects on pressure distribution and flow separation.**
This can be attributed to two main factors:
1)Increased Inertia:

At higher velocities, the airflow inertia becomes greater, making it more challenging to change direction (follow
the curved upper surface) at high AoA. This intensifies the adverse pressure gradient and promotes earlier flow
separation.

2)Boundary Layer Thickness:

As velocity increases, the boundary layer (the thin layer of slow-moving air adjacent to the airfoil) also thickens.
This thicker layer requires a stronger favorable pressure gradient (pressure decreasing in the flow direction) to
remain attached, which may not be achievable at high AoA.

C)Shockwave Formation (Case C9):

At high velocity (500 m/s, exceeding Mach 1), the airflow speed surpasses the local speed of sound. This
sudden change in speed creates a shockwave, a pressure discontinuity visible on the upper surface of the airfoil in
the pressure contour. The velocity contour depicts a localized region of high velocity followed by a sudden
decrease, both signatures of shockwave presence. Additionally, shock diamonds are visible downstream,
indicating the interaction of the shockwave with the flow and the resulting pressure and temperature fluctuations.

D)Connecting to Fluid Mechanics:

These observations directly link to established concepts in fluid mechanics:

1)Lift and Drag:

The pressure distribution across the airfoil influences both lift and drag. The pressure differential between the
upper and lower surfaces contributes to lift, while flow separation and increased skin friction due to a turbulent
boundary layer contribute to drag.

2)Boundary Layer Theory:

The behavior of the boundary layer plays a crucial role in flow separation. Understanding how the boundary
layer interacts with the pressure gradient is essential for predicting stall characteristics.

3)Compressible Flow:

When velocities approach or exceed the speed of sound, compressibility effects become significant.
Shockwave formation introduces additional complexities to the flow behavior and needs to be considered in high-
speed aerodynamics.

IV-B: COMPRESSIBILITY EFFECTS ON NACA 64A-010 AIRFOIL PERFORMANCE

The analysis of the NACA 64A-010 airfoil has provided valuable insights into the relationship between angle
of attack (AoA), velocity, and the resulting pressure distribution and flow behavior. However, for high-speed
flight regimes, it is crucial to consider the additional complexities introduced by compressibility
effects.Compressibility effects become important at high velocities. Below is the exploration of how
compressibility can impact the aerodynamic performance of the NACA 64A-010 airfoil.
A)Understanding Compressibility:

Compressibility signifies the inherent ability of a fluid to resist changes in density. At low velocities, air can be
approximated as an incompressible fluid, implying negligible variations in its density. However, as velocities
approach the speed of sound (Ma = 1), air molecules experience significant pressure fluctuations, leading to
noticeable density variations. These variations become increasingly prominent as the flow transitions into the
transonic and supersonic regimes. [3]

B)Impact on Aerodynamic Performance:

The introduction of compressibility significantly influences the aerodynamic performance of airfoils, primarily
through:

1)Lift Reduction:

At high velocities, shockwaves can form on the upper surface of the airfoil due to the localized flow exceeding
the speed of sound. These shockwaves disrupt the smooth airflow, potentially diminishing the crucial pressure
differential that generates lift.

2)Increased Drag:

Shockwave formation contributes to a rise in drag through two primary mechanisms:

i)Wave Drag:

Compressibility effects introduce challenges at high speeds. Supersonic flow struggles to smoothly follow an
airfoil's curvature, leading to shockwave formation. This rapid compression, while redirecting the flow,
dissipates energy. The conversion from supersonic to subsonic flow through the shockwave is not ideal,
resulting in wave drag – a force opposing the desired airflow. This loss of organized energy highlights the need
to consider compressibility and mitigate shockwaves for efficient airfoil design in high-speed regimes.

ii)Skin Friction Drag:

Within the realm of compressible flow, shockwaves present a double-edged sword for airfoil performance.
While crucial for redirecting supersonic flow to subsonic speeds, this process incurs an energetic penalty – wave
drag – due to the non-isentropic nature of the conversion. Furthermore, the interaction between the shockwave
and the boundary layer, the thin region of viscous flow adjacent to the airfoil, can significantly exacerbate skin
friction drag. The normally laminar boundary layer can be disrupted by the shockwave, triggering a transition to
turbulence. This turbulent flow presents a rougher interface to the airflow compared to its laminar counterpart,
resulting in increased frictional resistance between the air and the airfoil surface. In essence, the shockwave acts
as a catalyst for flow transition, transforming the well-behaved laminar layer into a turbulent one, and
consequently augmenting skin friction drag. [4]

C)Critical Mach Number:

The critical Mach number (Ma_crit) serves as a critical parameter that defines the velocity at which
compressibility effects become significant. It represents the Mach number at which the local airflow velocity
reaches the speed of sound. The specific critical Mach number for the NACA 64A-010 airfoil depends on
factors such as its geometry and the prevailing flight conditions.
D)Case C9 (Ma = 1.46) and Shockwave Formation:

The analysis included Case C9, where the velocity reached 500 m/s, corresponding to a Mach number of 1.46
(assuming standard atmospheric conditions). This velocity surpasses the critical Mach number, and the observed
shockwave formation on the upper surface of the airfoil confirms this phenomenon. The pressure distribution,
characterized by a sharp transition, and the velocity contour, depicting a localized high-velocity zone followed
by a decrease, are all signatures of a shockwave's presence.

E)Conclusion:

Compressibility effects emerge as a vital consideration in airfoil performance at high velocities. A thorough
understanding of how compressibility impacts lift, drag, and shockwave formation is paramount for designing
airfoils that operate efficiently in the transonic and supersonic regimes. Future investigations could explore the
influence of airfoil design features on mitigating compressibility effects and optimizing performance at high
speeds.

IV-C: DESIGN CHOICES, PERFORMANCE IMPACT, AND STUDY CONSIDERATIONS: A HOLISTIC ANALYSIS

My initial simulations on the NACA 64A-010 airfoil yielded results that seemed unrealistic at high velocities
(100-200 m/s). Consulting various resources on airfoil performance proved invaluable. These resources
highlighted the crucial link between airfoil design and operational velocity range. [5]

The NACA 64A-010, unlike airfoils designed for very low speeds, finds application in various fields, including
large wind turbines. While it can operate at a wider range of velocities, for accurate results in wind turbine
simulations, it's essential to consider the appropriate operating regime. Based on my research and comparisons
with established reports, the NACA 64A-010 seems to perform best within the ideal gas regime at velocities
relevant to wind turbine operations (typically ranging from 5 m/s to 30 m/s or even higher depending on specific
locations). [2]

The simulations conducted at these lower velocities (10-30 m/s) with varying angles of attack yielded lift and
drag coefficients that closely matched those reported in other studies. This agreement validates the approach and
reinforces the importance of selecting appropriate operating conditions for the chosen airfoil geometry,
especially when considering applications like wind turbines.

The analysis of the NACA 64A-010 airfoil highlights the influence of design choices on its aerodynamic
performance. Here's a breakdown of how specific features and the inherent trade-offs impact its behavior:

A)Design Feature: Airfoil Thickness

Strengths:

The NACA 64A-010 is a relatively thick airfoil (compared to airfoils designed for high speeds). This thickness
contributes to:

1)Higher Lift Coefficients:

Thicker airfoils generally generate higher lift coefficients at low angles of attack due to the larger camber
(curvature) of the upper surface. This makes the NACA 64A-010 suitable for applications requiring good low-
speed lift, such as take-off and landing of airplanes or in large wind turbines for which it is primarily ued .
Limitations:

1)Increased Drag:

Thicker airfoils also experience higher drag coefficients compared to thinner airfoils. This is because the thicker
profile creates a larger surface area for air to overcome, resulting in increased skin friction drag.

2)Susceptibility to Stalling:

As the angle of attack increases, the thicker profile can lead to a more prominent adverse pressure gradient on
the upper surface. This can cause earlier flow separation and stall at higher AoA compared to thinner airfoils.

B)Limitations of the Study:

1)Limited Velocity Range:

The simulations primarily focused on velocities between 10 m/s and 30 m/s. While relevant for wind turbine
applications, this range might not capture the full performance spectrum of the NACA 64A-010. Extending the
analysis to higher velocities (potentially exceeding 30 m/s) would have been necessary to explore
compressibility effects that become significant at transonic speeds.

2)Simplified Turbulence Modeling:

The analysis lincorporated some level of viscosity modeling to account for friction between air and the airfoil.
However, it relied on simplified approaches. Real-world airflow exhibits complex viscous phenomena,
particularly at high angles of attack and under varying wind conditions. Employing more sophisticated
turbulence models could provide a more nuanced understanding of boundary layer behavior and flow separation,
leading to more accurate predictions of lift, drag, and stall characteristics.

3)Single Airfoil Geometry:

The analysis focused solely on the NACA 64A-010 airfoil. While it offers valuable insights into this specific
design, it doesn't account for the vast array of airfoil geometries used in wind turbines. Investigating the
performance of other airfoil shapes, potentially with modifications to camber or thickness for the NACA 64A-
010 itself, could offer a broader perspective on wind turbine optimization.

4)Computational Constraints:

Depending on the computational resources available, the simulations involved limitations in mesh resolution or
turbulence model complexity. These limitations can affect the accuracy of the results, particularly for high
angles of attack or complex flow phenomena.

5)Validation with Experimental Data:

While the results aligned with established reports, the analysis would benefit from validation with experimental
data from wind tunnel tests on the NACA 64A-10 airfoil specifically. This would provide a stronger foundation
for the computational results and enhance confidence in the predicted performance.
REFERENCES

[1] [Online]. Available: https://www.simscale.com/blog/turbulence-cfd-analysis/.


[2] [Online]. Available:
https://ntrs.nasa.gov/api/citations/19880003923/downloads/19880003923.pdf.
[3] [Online]. Available: https://www.naa.edu/airfoil-
design/#:~:text=An%20airfoil%20(or%20aerofoil%20in,airfoil-shaped%20cross-sections..
[4] [Online]. Available:
https://www.narcis.nl/publication/RecordID/oai%3Atudelft.nl%3Auuid%3Adff7d6b3-b00a-
4b22-8dac-
eaf1d963722e/uquery/jet%20OR%20turbine%20OR%20turbo%20OR%20turbofan%20OR%2
0turbojet%20AND%20aerofoil%20OR%20aerofoils%20OR%20airfoil%20OR%20airfoils%2
0OR%20blade%.
[5] [Online]. Available:
https://chem.libretexts.org/Bookshelves/Introductory_Chemistry/Introductory_Chemistry_(CK
-12)/14%3A_The_Behavior_of_Gases/14.01%3A_Compressibility.
[6] [Online]. Available: https://skybrary.aero/articles/friction-drag.
[7] [Online]. Available: https://www.researchgate.net/figure/Lift-and-moment-coefficient-loops-
for-a-NACA-64A010-airfoil-at-Mach-number-0796_fig4_327629358.

You might also like