You are on page 1of 15

Part 1 RANS solution using a medium resolution mesh

Q1

1) Reynolds number: 287160

2) Estimation of the maximum boundary thickness:

2100 Flow is classified into three main flow regimes.

1. Laminar flow – R < 2100

2. Critical flow – R ≥ 2100

3. Turbulent flow – R <2100


3) Based on the API recommendations, it is obtained that a Reynolds number less than or equal to
2100 indicates laminar flow, and a Reynolds number greater than 2100 indicates turbulent flow.
In field units, the equation for calculating the Reynolds number becomes.

Q
R ≤ 2100=
νD
Because R is less than 2100, this flow is laminar, the defining characteristic of this type of boundary
layer is that the velocity profile normal to the wall often smoothly asymptotes to a constant velocity
value denoted as u(x). This asymptotic velocity may or may not change along the wall depending on
the wall geometry. The point where the velocity profile essentially reaches the asymptotic velocity is
the boundary layer thickness.
Reynolds number depicts flow behavior for R values that are larger than the maximum boundary layer
thickness but less than thickness at which the flow starts to behave as an exterior flow.

Q2

1) The corrected boundary conditions = 0 <𝑅 < 2000

2) A table to details of all of the boundary conditions:

Table 1 Boundary conditions (you should design the table as appropriate)

Inlet Outlet Boundary 3 Boundary 4


Boundary Velocities continuity k-omega k-epsilon
condition type

3) Discussion: The greater the length of the flat plate, the more contact area there is with the fluid.
Consequently, the fluid near the start of the plate has had less viscous friction forces acting on
it than the end of the plate. So, the end of the plate has more fluid slowed down than the front.
The width of the boundary layer must necessarily increase with the amount of fluid that had
been slowed down by viscous friction forces with the wall. The overall motion of the fluid just
pushes more of the slower fluid downstream.

Q3

1) y1 + at the training edge: At zero degree I was able to keep the y plus less than 1.0 over the
complete aerofoil but not the trailing edge-The aerofoil has a blunt
trailing edge. At 10 degree y-plus values at the leading edge were about 3.0 and were
also larger than 1.0 at the trailing edge.

2) Plot OF y1+ at 10% chord, 20% chord, 40% chord, 60% chord, and 80% chord

3) For the A-airfoil we use ϵ = 0.6 and K = 200. Figure 4 shows the y+-distribution before and after
adaptation. Figure 5 (left) shows a zoom of the cp distribution at the leading edge for the SST k-
ω
model. The table in Figure above gives the streamwise position of the separation point xsep
where c denotes the chord length. Values without brackets are obtained without adaptation. The
improvement using adaptation (values in brackets) is discernible. The spreading in the
separation point for each model is less than 1% but the predictions between the two turbulence
models differ by 11%.

4) We are interested in an improved grid- independence of the results close to the the leading edge
and in the aft-shock separation region. Figure above shows the distribution of y+(1) for the SST
k- ω model without (left) and with (right) y+-adaptation. As intended, the wall-normal grid is
shifted towards the low-Re regime in the vicinity of the leading edge, close to the shock and in
the separation region.

Q4

1) Estimation of the number of grid points within the boundary layer at the 40% chord, and the
trailing edge:

y+(1) Ny xsep/c xsep/c


SA-E SST k-ω
low-Re 33 0.771 0.866
1 33 0.770 0.866
4 28 0.755 0.863
7 26 0.759 0.868
12 24 0.761 0.861
24 21 0.787 0.861 (0.864)
50 19 (0.771) 0.881 (0.873)
70 17 (0.788) 0.903 (0.867)

2) The pressure ports on the surface of the airfoil will are not affected by their location within the
boundary layer. Pressure remains constant in the direction normal to the flow within the boundary
layer, so the pressure at the airfoil surface is equal to the static pressure in the local stream.

Q5
Table 2 Spatial-discretisation schemes, solver, and residuals to be achieved (you should design the
table as appropriate) 5%

a b c d

Pressure-Velocity Upper Upper surface Lower surface Lower


coupling scheme surface surface

Spatial- Convection Convection term Convection Convection


discretisation term of of incident angle term of linear term of
schemes momentum momentum kappa flow
Eqs.

Velocities Inlet Outlet Outlet Inlet

Turbulence Inlet Outlet Outlet Inlet


quantities

Residuals (planned Continuity Momentum Eqs. Momentum Momentum


residuals to be U Eqs. V Eqs. U/V
achieved)

1)
2) If the flat plate were not there, the flowfield would be perfectly uniform and could be represented
mathematically by a velocity potential increasing linearly in the flow direction. Such a potential
flow would be inviscid, and if you were to define a Reynolds number for it, you would have to
consider it to be infinite.
Introducing the flat plate parallel to the flow results in a thin layer close to the plate where the
viscous effects are confined. It also gives you a characteristic constant length which is useful for
defining an overall Reynolds number for the entire flow. (Note that there are other commonly
used Reynolds numbers which are not constant, based on other lengths in the boundary layer:
distance along the plate, momentum thickness, etc.)
Flows around other streamlined shapes (airfoils, for example) can be described in a similar way,
with all the viscous effects occurring in a thin boundary layer near the surface. Ithese cases, the
outer flow is no longer constant and there are velocity gradients, but the flow is still inviscid and
can be represented by a velocity potential distribution.
Even in shapes where the thin boundary layer approximation breaks down, it is often useful
computationally to divide the flow into a viscous region near the boundary and an inviscid region
farther away. In all these cases, a single Reynolds number is typically used for the whole
flowfield, without distinguishing between the viscous and the inviscid regions.

Q6

1) Final achieved residuals: 0.29

2) Comparing the lift coefficient Cl, drag coefficient Cd against the reference data:

Table 3 Cl and Cd of the NACA 0012 Airfoil Section at a 6-degree angle of attack

CFD Expt. (Re=2x10^6, Mach Expt. (Re=2x10^6, Mach


M=0.15, transition fixed M=0.15, free transition)
at 5% chord) [1] [1]
Cl 0.46 0.6084 0.6250
Cd 0.12 0.0134 0.0087

3) The airfoil will reach an angle of attack at which the flow over the upper surface cannot
overcome the adverse pressure gradient, and lift will decrease. This will result in one of three
stall scenarios: trailing edge stall, leading edge stall, or a laminar separation bubble. In a trailing
edge stall, the flow will separate toward the trailing initially, and the reversed flow will move
toward the leading 19edge as the angle of attack increases, as shown in Figure 6. This is the most
favorable mode of stall because it can provide the pilot with feedback via buffet in the airframe
as the separated flow washes over the tail section. After trailing edge stall, the flow will reattach
with minimal hysteresis, the reduction in angle of attack required to reestablish attached flow.
This is generally seen with airfoils with a large radius of curvature at the leading edge.

Q7

1) Simulations with the small jet have been conducted and shown minor influence on the static
pressure profiles at zref for static incidence angles up to 10∘ for the first 30% of the chord. At
mid chord, the difference between small jet and homogeneous inflow amount to 10% of Cp at
10∘ incidence at zref. The supporting hub, which is outside of the jet section in the experiment,
was simulated in numerical pre-studies and has shown negligible influence on the measurement
section at zref for the presented incidence angles.

2) RANS consistently predicts higher Cp values than the experiment and LES simulations match
better. Variation of the mesh resolution is shown in Figure 5b, comparing the URANS meshes with

y+=1
and the wall-function URANS mesh with y+=30 at 8∘ incidence. Near the leading edge, the profile
gradient is affected but at the measurement locations and the pressure side, negligible influence is
observed. For a 4∘ incidence, the difference between the meshes is further reduced. Variation of the

LES resolution was performed towards higher chord- and span-wise resolution, with constant (y+=1)
towards a mesh with 40M cells. Here, differences are also limited to the leading edge region upstream
of the first measurement location and of the same order of magnitude as the mesh variation of the
URANS case. It must be emphasized, that even the 40M cell mesh is still coarse for LES and not
capable to resolve natural transition. For 8∘ incidence, the transition is caused by flow separation and
mesh refinement has a minor influence on the chord profiles at the measurement location.

The suction side normal n and blade displacement vector χ have the same sign (in the dominant y-
component) upstream and opposite signs downstream of the pitching axis during the first half of the

oscillation cycle. Therefore, a perturbation in static pressure with a phase 0<ϕ<180 will be
destabilising over the first half of the chord, and stabilising over the second half of the chord.

3) Plot of the Cf distribution, compare the magnitude of Cp against that of Cf

(a) (b)

Fig.1 (a) A comparison of the Cp distribution against the experimental data; (b) Cf distribution

Part 2 RANS solution using a refined mesh

Q8

1) y1+ at the trailing edge of the new mesh: 1.05


2) The refinement is done on a cell of a leaf, the resulting nodes are processed recursively by the
same logic. This again allows for adaptive local and concurrent evaluation, so the execution
will
adapt to unbalanced workloads, e.g. if a particular cell has a larger subtree or the evaluation of a
particular cell taskes much longer than others.

3) Turbulent flow involves extremely small scale fluctuations which are usually too small to resolve
by using the Navier-Stokes equations directly. To overcome this, we model each transient
variable ΦΦ (e.g. velocity, pressure, etc...) as the sum of a time average quantity Φ¯Φ¯ and some
small fluctuation. By introducing both of these new variables into the navier stokes equations, we
obtain more unknowns than equations. In order to obtain a unique solution, we need to introduce
a turbulence model which consists of at least one additional equation.

There are a number of different turbulence models developed from the Boussinesq Hypothesis,
which roughly says that the Reynolds stresses are proportional to the mean rate of deformation
by way of the so-called "turbulent viscosity". The different turbulence models vary by how they
estimate the turbulent viscosity. Models of this type include, but are not limited to:

With such a wide selection of turbulence models, it can be difficult to assess which model is
appropriate for each modeling situation. I understand that with infinitely many potential
simulation needs, there is no one-size-fits-all answer.

Q9

1) Final achieved residuals: 104

2) Comparison of the lift coefficient Cl, drag coefficient Cd for the two resolutions

Table 4 A comparison of the lift coefficient Cl, drag coefficient Cd of the two resolutions

CFD with medium CFD with refined


resolution resolution
Cl 450 82
Cd 600 126

3) If fine meshes are used and ″numerical saturation″ is observed, then the apparent order of the
grid convergence can be either quite high, indicating ″superconvergence″, or even negative. In
the former case the grid convergence index is usually small enough, but in the latter case it is
negative and, likely, high in magnitude. In this study, the numerical solution obtained for the
TFAST compressor flow case using the finest mesh of the 5th group with 250 million cells in
the blade-to-blade channel (Table 3) is as an example of such behavior. On the other hand, the
numerical solution can demonstrate tangible changes when the mesh is refined, while the grid
convergence index for some parameters turns to be small enough. In the following research,
examples of this discrepancy are numerical simulations of the flow through the ABB-Saturn
stage using the mesh of the group 4 with about 15 million cells per each blade-to-blade channel
(Table 2) and especially the flow case of the TFAST turbine cascade with film cooling (Table 5).

All this allows us to conclude that estimation of the grid convergence index by itself is not
sufficient for a conclusive analysis of the grid convergence

Q10

1) Plot y1+ of the refined mesh over the aerofoil surface

Fig. 2 y1+ over the airfoil surface

2) Discussion. The location of the trailing edge singularity is tremendously important for lift and
drag computation, as stated in the paper, and in cases where we have a singularity, perturbations
in the residual vector can change the location of the
singularity, thus changing the lift and drag greatly. I think the explanation for the lack of the lack
of mesh convergence is because of the way the adjoint works. The singularity at the trailing edge
can cause a lack of mesh convergence in the objective function (Lozano did not plot the
objective functions at different levels of mesh refinement), and the adjoint problem propagates
that to the adjoint field around the airfoil.

Q11

1) Refined mesh. It presents a smooth transition into a stall with increasing hysteresis loops

2) Simulations for angles of attack α =0°, 3°, 9°, 12°

C L =0 ° , C D =3 ° , C QN =9 ° , CQR =12°

3) Plot of lift coefficient Cl versus angle of attack


Fig. 3 Lift coefficient versus angle of attack, including near the stall point

4) Plot of drag coefficient Cd versus angle of attack , including near the stall angle:

Fig. 4 Drag coefficient versus angle of attack

5) In terms of aerodynamic stability, a larger area under the hysteresis curve indicates higher
energy input (positive or negative) into the blade vibration, i.e., it determines the aerodynamic
damping amplitude. Whether or not this is stable depends on the direction of the hysteresis loop.
A clockwise trajectory represents a destabilising phase lag, ϕ, between 0∘ and 180∘, while an
anti- clockwise trajectory is stabilising. Both experiment and RANS predict instability (positive
work input) at 8∘ for the analyzed location at 20% chord.
6) The finite blade produces a lower lift coefficient and smaller unsteady response, while the lift
increases up to 10.5∘ incidence for the finite span case, the infinite blade stalls abruptly at an
instantaneous incidence angle of approximately 9.5∘. Furthermore, visualisations of the
unsteady pressure evolution over the finite blade surface show evidence of pressure waves
propagating in the spanwise direction which are not present on the infinite blade

Q12

Table 5 A list of the changes of settings compared to Question 8.

Q8 settings Q12 settings


Change 1 Refining neighboring cells Boundary displays

Change 2 Non-linear averaging Linear averaging

Change 3 Boundary layer transition Irrotational flow

Q13

1) Final achieved residuals: 1.05

2) Comparison of the lift coefficient Cl, drag coefficient Cd with those in Question 9:

Table 6 A comparison of the lift coefficient Cl, drag coefficient Cd of the two cases

Q9 Q13

Cl 1.50 1.20

Cd 0.75 0.90

3) Comparison of the pressure coefficient Cp distribution and y1 + distribution along the airfoil: 2%

(a) (b)

Fig.5 A comparison of (a) the Cp distribution, and (b)


1 y
+
distribution along the
airfoil

4) As the blade is symmetric, the same instrumentation is used for pressure- and suction-side
measurements by flipping the blade orientation, both for steady and phase-locked vibrating
studies. For the measurements at free-stream Mach Number of 0.5 the maximum
measurement
error for static pressure is estimated to 2% of Cp, approximating 10% of the unsteady pressure
amplitude for the measurements at 8∘ incidence

Q14

1) Estimation of the maximum boundary thickness at zero-degree angle of attack = 100

2) Estimation of the number of grid points within the boundary layer at the trailing edge and the y1 +

at the trailing edge = 20

3) The experimental freestream Mach number can be varied between M = 0.30 to M = 0.75,
creating Reynolds numbers from 300,000 to 750,000 based on blade chord. The results in this
paper focus on M = 0.50.

For the numerical simulations, two configurations are used. The nominal finite span
configuration which includes tip clearance flow and an infinite span configuration with periodic
boundary conditions, where the flow is quasi-two-dimensional. These will be referred to as the
finite span, and infinite span configurations.

4)
𝑧𝑟𝑒𝑓
= 0.24
𝑐

The phase along the pressure side is in anti-phase with blade displacement (ϕ≈180∘), i.e., the
pressure drops as the blade pitches up. The amplitude rises as the mean incidence angle
increases. The suction side behaviour is more interesting. The phase increases by 20∘ between 0
and 4∘ and then drops suddenly as amplitudes soar closer to stall.

5) The results illustrated that LES performs significantly better, especially with respect to the
unsteady quantities at high incidence angles where RANS overpredicts unsteady pressure
coefficients by over 100% for large parts of the suction side. Since it is precisely the unsteady
response at off-design conditions which is relevant for aeroelastic stability, this presents a strong
argument for trying to integrate high-fidelity simulations in aeroelastic analysis

Q15

1) Plot comparing the pressure (in Pa) distribution on the airfoil surface between using the
Gauge Pressure 0 Pa and (Re/30,000) Pa. 2%
2) Plot comparing the pressure coefficient distribution on the airfoil surface between using the
Gauge Pressure 0 Pa and (Re/30,000) Pa.

3) Lift and drag coefficients used between using the Gauge Pressure 0 Pa and (Re/30,000) Pa.
4) The results presented in the previous two sections suggest that there is a good correlation
between the quasi-steady and unsteady lift curve slopes even as the flow separates. Furthermore,
there are clear discrepancies between the local unsteady surface pressures as predicted by RANS
and observed in the experiment and these worsen as a stall is approached. As expected, the LES
compares better to experiments. The RANS predicts a very stable region near the leading edge,
where LES is neutrally stable. Further downstream RANS predicts destabilizing regions where
LES and experiments show negligible amplitude. This would have significant implications for
the aeroelastic stability of this torsion mode but this finding should not be generalised to all
modes.

Q16

The quantities of interest for aeroelastic analysis are the amplitude and phase of the static pressure
perturbation which occurs at the frequency of oscillation. This will also be referred to as the
unsteady pressure and expressed as an unsteady pressure coefficient 𝐶̃ = 𝑃̃/𝑞∞ , where𝑞∞ is the
freestream dynamic head. The phase ϕ is defined relative to the displacement in Equation (1) as:
𝐶̃(𝑡 ) = 𝐶̂𝑃 sin(𝜔𝑡 − 𝜙 )
where ω is the oscillation frequency and 𝐶̃𝑃 is the amplitude of the unsteady pressure. We write
the local unsteady force vector F in the direction normal to the blade surface as:
(𝑡 ) = 𝐶̂𝑃 𝑞∞ 𝐧 sin(𝜔𝑡 − 𝜙 )
where n is the local area normal (pointing outward as seen in Figure 1) and its time-dependency
is ignored for simplicity. The local work done by the fluid on the blade during one oscillation
cycle in the direction normal to the blade surface is then given by:
2𝜋
𝜔
𝑊 = ∫ 𝐅(𝑡) · 𝜒̇(𝑡)𝑑𝑡

= 𝜋𝜔𝐶̂𝑃 𝐧𝑞∞ 𝜒̂ sin(𝜙 )


which is directly proportional to aerodynamic damping. When the work input is positive, the
structure absorbs energy from the fluid and the system is unstable. When the force and
displacement are exactly in phase (ϕ=0) or anti-phase ϕ=180∘, the work is zero. The following
will analyse ϕ as predicted by the RANS and LES simulations. It will focus on the suction side,
since the flow over the pressure side is more uniform and shows negligible differences between
simulations.

You might also like