scispace - formally typeset
Open AccessJournal ArticleDOI

Direct numerical simulations of forced and unforced separation bubbles on an airfoil at incidence

Reads0
Chats0
TLDR
In this paper, the authors present a direct numerical simulation of laminar separation bubbles on a NACA-0012 airfoil at Re-c = 5 x 10(4) and incidence 5 degrees.
Abstract
Direct numerical simulations (DNS) of laminar separation bubbles on a NACA-0012 airfoil at Re-c = 5 x 10(4) and incidence 5 degrees are presented. Initially volume forcing is introduced in order to promote transition to turbulence. After obtaining sufficient data from this forced case, the explicitly added disturbances are removed and the simulation run further. With no forcing the turbulence is observed to self-sustain, with increased turbulence intensity in the reattachment region. A comparison of the forced and unforced cases shows that the forcing improves the aerodynamic performance whilst requiring little energy input. Classical linear stability analysis is performed upon the time-averaged flow field; however no absolute instability is observed that could explain the presence of self-sustaining turbulence. Finally, a series of simplified DNS are presented that illustrate a three-dimensional absolute instability of the two-dimensional vortex shedding that occurs naturally. Three-dimensional perturbations are amplified in the braid region of developing vortices, and subsequently convected upstream by local regions of reverse flow, within which the upstream velocity magnitude greatly exceeds that of the time-average. The perturbations are convected into the braid region of the next developing vortex, where they are amplified further, hence the cycle repeats with increasing amplitude. The fact that this transition process is independent of upstream disturbances has implications for modelling separation bubbles.

read more

Content maybe subject to copyright    Report

J. Fluid Mech. (2008), vol. 602, pp. 175–207.
c
2008 Cambridge University Press
doi:10.1017/S0022112008000864 Printed in the United Kingdom
175
Direct numerical simulations of forced
and unforced separation bubbles on an
airfoil at incidence
L. E. JONES, R. D. SANDBERG AND N. D. SANDHAM
Aerodynamics and Flight Mechanics Research Group, School of Engineering Sciences, University of
Southampton, Southampton, SO17 1BJ, UK
(Received 21 August 2007 and in revised form 23 January 2008)
Direct numerical simulations (DNS) of laminar separation bubbles on a NACA-0012
airfoil at Re
c
=5× 10
4
and incidence 5
are presented. Initially volume forcing is
introduced in order to promote transition to turbulence. After obtaining sufficient
data from this forced case, the explicitly added disturbances are removed and the
simulation run further. With no forcing the turbulence is observed to self-sustain, with
increased turbulence intensity in the reattachment region. A comparison of the forced
and unforced cases shows that the forcing improves the aerodynamic performance
whilst requiring little energy input. Classical linear stability analysis is performed
upon the time-averaged flow field; however no absolute instability is observed that
could explain the presence of self-sustaining turbulence. Finally, a series of simplified
DNS are presented that illustrate a three-dimensional absolute instability of the two-
dimensional vortex shedding that occurs naturally. Three-dimensional perturbations
are amplified in the braid region of developing vortices, and subsequently convected
upstream by local regions of reverse flow, within which the upstream velocity
magnitude greatly exceeds that of the time-average. The perturbations are convected
into the braid region of the next developing vortex, where they are amplified further,
hence the cycle repeats with increasing amplitude. The fact that this transition process
is independent of upstream disturbances has implications for modelling separation
bubbles.
1. Introduction
Under an adverse pressure gradient a boundary layer may separate, leading to
reverse (upstream) fluid flow. Within the separated region disturbances are strongly
amplified, typically leading to transition to turbulence. The resultant turbulent flow
enhances mixing and momentum transfer in the wall-normal direction, and causes
the boundary layer to reattach. This system of laminar separation, transition and
turbulent reattachment is referred to as a laminar (or transitional) separation bubble
(LSB), and is typically associated with flows at low to moderate Reynolds numbers.
When present on an airfoil, laminar separation bubbles have a marked effect upon
aerodynamic performance. Drag forces are typically increased, and the presence of
a separation bubble may determine stall behaviour (Gault 1957). The phenomenon
of bubble bursting, where a small increase in incidence leads to a sudden increase in
bubble length, causes a dramatic loss in aerodynamic performance and hence is an
important consideration in low-Reynolds-number airfoil design.

176 L. E. Jones, R. D. Sandberg and N. D. Sandham
Using the results of Gaster (1967), Horton (1969) was the first to describe the
time-averaged structure of a laminar separation bubble, and proposed an empirical
model for predicting bubble behaviour. Despite refinements such as the use of the
e
n
transition prediction method, modelling of low-Reynolds-number effects and the
dependence on background turbulence levels, present day models do not adequately
predict bubble bursting or unsteady behaviour.
More recently, advances in understanding of laminar separation bubbles have been
made via numerical methods. The first numerical simulations of separation bubbles
were limited either to two-dimensional analysis (Pauley, Moin & Reynolds 1990), or
only studied primary/linear instability and did not resolve transition (Pauley 1994;
Rist 1994). Being less computationally expensive, linear stability analysis could be
performed before fully resolved direct numerical simulations (DNS) were possible.
Hammond & Redekopp (1998) performed local analysis of separated boundary layer
profiles in order to determine whether absolute instability could be observed, as it
had been for separated shear layers (Huerre & Monkewitz 1985) and bluff-body
wakes (Hannemann & Oertel 1989). For certain profiles local absolute instability
was observed, depending on both the maximum reverse flow and the height of
the reverse flow region. Similar criteria were investigated and confirmed later by
Rist & Maucher (2002). Hammond & Redekopp (1998) found that for profiles at
Re
δ
=10
3
, a minimum reverse flow velocity of 20 % was required to observe local
absolute instability. Theofilis (2000) performed global linear stability analysis of a
laminar separation bubble, specifying a spanwise wavenumber, β, and computing the
resultant two-dimensional disturbance eigenvectors. A temporally growing stationary
mode was observed, associated with unsteadiness at the reattachment point but not
affecting the separation point. Growth rates were found to be significantly lower than
those associated with amplification of convective instabilities within the shear layer;
however Theofilis (2003) suggests that the existence of this mode may potentially be
relevant to the phenomenon of vortex shedding from separation bubbles as observed
by Pauley et al. (1990).
The first numerical simulations to fully resolve transition to turbulence within a
laminar separation bubble were conducted by Alam & Sandham (2000) and Spalart
& Strelets (2000). Alam & Sandham (2000) performed DNS of a laminar separation
bubble on a flat surface, induced by upper boundary transpiration. Performing linear
stability analysis on analytic velocity profiles similar to those observed in the DNS,
Alam & Sandham found that reverse flow greater than 15 % would be required
in order to sustain absolute instability, compared to an observed reverse flow of
only 4–8 %. As a result, it was concluded that the transition process was driven
by convective instability. Spalart & Strelets (2000) conducted DNS of a laminar
separation bubble for the purpose of assessing turbulence models. No unsteadiness
was introduced and inflow disturbances were less than 0.1 % of the free-stream
velocity; however transition to turbulence was still observed. As a result the study
stated that entry-region disturbances (referring to Tollmien–Schlichting, or TS, type
waves) could be discarded as the mechanism behind transition; however the study
also stated that the magnitude of reverse flow present was unlikely to be sufficient
to sustain absolute instability. In the study of Spalart & Strelets three-dimensionality
was observed to occur rapidly, with no clear regions of primary or secondary
instability, whereas Alam & Sandham observed -vortex-induced breakdown. Hence
the first two fully resolved DNS of laminar separation bubbles apparently observed
different instability mechanisms leading to transition, and different transitional
behaviour.

Forced and unforced separation bubbles on an airfoil at incidence 177
Marxen et al. (2003) performed a combined DNS and experimental study of an LSB
on a flat plate. Periodic two-dimensional disturbances were introduced upstream of
separation, and three-dimensionality was introduced via a spanwise array of spacers.
The separated shear layer was observed to roll up to form vortices which subsequently
broke down to turbulence. The same configuration was studied further by Lang, Rist
& Wagner (2004) and again by Marxen, Rist & Wagner (2004) in order to quantify the
respective roles of two-dimensional and three-dimensional disturbances. Marxen et al.
concluded that transition was driven by convective amplification of a two-dimensional
TS wave, which also determined the length of the bubble, and that the dominant
mechanism behind transition is an absolute secondary instability in a manner first
proposed by Maucher, Rist & Wagner (1997), and investigated further by Maucher,
Rist & Wagner (1998).
It is clear that numerical simulations of separation bubbles can differ markedly in
behaviour. Unlike wake and shear layer flows, the presence of regions of absolute
instability within laminar separation bubbles has not yet been confirmed by linear
stability analysis of DNS or experimental velocity profiles, and stability characteristics
of separation bubbles not well defined in all cases. With continued advances in
computing power, it is now possible to perform DNS of laminar separation bubbles
on full airfoil configurations. This contrasts with previous numerical studies, which
have been limited to bubbles on flat plates or other simplified geometries in order to
reduce the computational cost. The advantage of studying full airfoil configurations
is that the bubble can interact more strongly with the potential flow, in particular via
the Kutta condition at the trailing edge. The bubble will be closer in nature to those
observed under flight conditions, and the influence of the bubble behaviour on the
aerodynamic performance of the airfoil can be observed directly.
The purpose of the current study is to investigate the dependence of bubble
behaviour on the presence of boundary layer disturbances, and to investigate the
role of instability mechanisms in separation bubble transition. First, data from both
forced and unforced three-dimensional simulations of a laminar separation bubble
on a NACA-0012 airfoil will be compared. Classical linear stability analysis will then
be performed upon the time-averaged flow fields obtained from the DNS, in order
to determine whether any local absolute instability is present. Finally, a series of
computationally inexpensive simulations will be presented, intended to explain the
self-sustained transition to turbulence observed in the first part of the study.
2. Direct numerical simulations
2.1. Simulation geometry
The chosen airfoil geometry is a NACA-0012, extended to include a sharp trailing
edge and rescaled to unit chord length. The coordinate system for the curvilinear
C-type grids used in all simulations is given in figure 1. Grids are equidistantly spaced
in the z-direction for three-dimensional simulations. The two parameters governing
domain size are the wake length W , and the domain radius R. The airfoil chord is
used as the reference length scale and the coordinate system is defined such that the
trailing edge is located at (x, y)=(1, 0).
2.2. Governing equations
All simulations were run at a Reynolds number based on airfoil chord of Re
c
=5×10
4
,
and Mach number M =0.4 unless otherwise stated. A compressible-flow formulation
was chosen so that sound waves originating at the trailing edge could also be studied

178 L. E. Jones, R. D. Sandberg and N. D. Sandham
R
R
W
ξ
η
Figure 1. Topology of the computational domain.
(Sandberg, Sandham & Joseph 2007). The DNS code directly solves the unsteady,
compressible Navier–Stokes equations, written in curvilinear form as
Q
∂t
+
E
∂ξ
+
F
∂η
+
G
∂z
=
R
∂ξ
+
S
∂η
+
T
∂z
. (2.1)
The conservative vector Q, inviscid flux vectors E, F and G, and the viscous vector
terms R, S and T are defined as
Q =
ρ
ρu
ρv
ρw
E
t
, E =
ρU
ρuU +
x
ρvU +
y
ρwU
(E
t
+ p)U
, (2.2)
F =
ρV
ρuV +
x
ρvV +
y
ρwV
(E
t
+ p)V
, G =
ρw
ρuw
ρvw
ρww + p
(E
t
+ p)w
, (2.3)
R =
0
τ
xx
ξ
x
+ τ
xy
ξ
y
τ
yx
ξ
x
+ τ
yy
ξ
y
τ
zx
ξ
x
+ τ
zy
ξ
y
Q
x
ξ
x
+ Q
y
ξ
y
, S =
0
τ
xx
η
x
+ τ
xy
η
y
τ
yx
η
x
+ τ
yy
η
y
τ
zx
η
x
+ τ
zy
η
y
Q
x
η
x
+ Q
y
η
y
, T =
0
τ
xz
τ
yz
τ
zz
Q
z
, (2.4)
where ρ is the fluid density, u, v and w are velocity components in the Cartesian x, y
and z directions, p is the pressure, and E
t
is the total energy per unit volume, defined
as
E
t
= ρe +
1
2
ρ(uu + vv + ww), (2.5)
where
e =
T
γ (γ 1)M
2
. (2.6)

Forced and unforced separation bubbles on an airfoil at incidence 179
Metric terms are defined as
ξ
x
=
y
η
J
x
=
y
ξ
J
y
=
x
η
J
y
=
x
ξ
J
, (2.7)
noting that terms ξ
z
and η
z
are both equal to zero for computational grids with no
spanwise variation, as used in the current study, and the Jacobian J is defined as
J = x
ξ
y
η
x
η
y
ξ
. (2.8)
The contravariant velocities U and V are defined as
U = ξ
x
u + ξ
y
v, V = η
y
v + η
x
u, (2.9)
and the stress terms τ
ij
as
τ
ij
=
µ
Re
∂u
i
∂x
j
+
∂u
j
∂x
i
2
3
µ
Re
∂u
k
∂x
k
δ
ij
. (2.10)
The terms Q
i
comprise the conduction and work terms of the energy equation,
Q
i
= q
i
+ u
j
τ
ij
, (2.11)
where
q
i
=
µ
(γ 1)M
2
RePr
∂T
∂x
i
. (2.12)
Viscosity is calculated using Sutherland’s law,
µ = T
3/2
1+C
T + C
,C=0.368
˙
6, (2.13)
and finally, the perfect gas law relates p, ρ and T
p =
ρT
γM
2
. (2.14)
The primitive variables ρ,u,v and T have been non-dimensionalized by the free-
stream conditions and dimensionless parameters Re, Pr and M are defined using
free-stream (reference) flow properties. The ratio of specific heats is specified as
γ =1.4 and the Prandtl number as Pr =0.72.
2.3. Numerical method
Fourth-order-accurate central differences utilizing a five-point stencil are used for
spatial discretization. Fourth-order accuracy is extended to the domain boundaries by
use of a Carpenter boundary scheme (Carpenter, Nordstr
¨
om & Gottlieb 1999). No
artificial viscosity or filtering is used. Instead, stability is enhanced by entropy splitting
of the inviscid flux terms in combination with a Laplacian formulation of the viscous
terms (Sandham, Li & Yee 2002). The explicit fourth-order-accurate Runge–Kutta
scheme is used for time-stepping, and all cases were run with a constant time step of
t =1× 10
4
. Appropriate boundary conditions must be applied to avoid unphysical
reflections. At the free-stream (η
+
) boundary, where the only disturbances likely to
reach the boundary will be in the form of linear waves, an integral characteristic
boundary condition is applied (Sandhu & Sandham 1994). At the downstream exit
boundary (ξ
±
), which will be subject to the passage of nonlinear fluid structures,
a zonal characteristic boundary condition (Sandberg & Sandham 2006) is applied
for increased effectiveness. At the airfoil surface an adiabatic, no-slip condition is
applied.

Citations
More filters
Journal ArticleDOI

Global Linear Instability

TL;DR: A review of linear instability analysis of flows over or through complex 2D and 3D geometries is presented in this article, where the authors make a conscious effort to demystify both the tools currently utilized and the jargon employed to describe them, demonstrating the simplicity of the analysis.
Journal ArticleDOI

Dynamics of laminar separation bubbles at low-Reynolds-number aerofoils

TL;DR: In this paper, the laminar separation bubble on an SD7003 aerofoil at a Reynolds number Re = 66000 was investigated to determine the dominant frequencies of the transition process and the flapping of the bubble.
Journal ArticleDOI

Stability and receptivity characteristics of a laminar separation bubble on an aerofoil

TL;DR: In this paper, the authors investigated the aerofoil wake behind an NACA-0012 aerodynamic model and found that the wake exhibits a laminar separation bubble and no absolute instability in the classical sense.
Journal ArticleDOI

Acoustic and hydrodynamic analysis of the flow around an aerofoil with trailing-edge serrations

TL;DR: In this paper, a NACA-0012 aerofoil with and without serrations was modeled using an immersed boundary method to represent flat-plate trailing-edge extensions both without and with serrations.
Journal ArticleDOI

Direct numerical simulations of transition in a compressor cascade: the influence of free-stream turbulence

TL;DR: In this article, the flow through a compressor passage without and with incoming free-stream grid turbulence is simulated. And the authors reveal the mechanics of breakdown to turbulence on both surfaces of the blade, and three types of breakdowns are observed; they combine characteristics of natural and bypass transition.
References
More filters
Journal ArticleDOI

Compact finite difference schemes with spectral-like resolution

TL;DR: In this article, the authors present finite-difference schemes for the evaluation of first-order, second-order and higher-order derivatives yield improved representation of a range of scales and may be used on nonuniform meshes.
Book

Stability and Transition in Shear Flows

TL;DR: In this article, the authors present an approach to the Viscous Initial Value Problem with the objective of finding the optimal growth rate and the optimal response to the initial value problem.
Journal ArticleDOI

Direct simulation of a turbulent boundary layer up to R sub theta = 1410

TL;DR: In this paper, the turbulent boundary layer on a flat plate, with zero pressure gradient, is simulated numerically at four stations between R sub theta = 225 and R sub tta = 1410.
Journal ArticleDOI

Viscous-inviscid analysis of transonic and low Reynolds number airfoils

TL;DR: In this article, a method of accurately calculating transonic and low Reynolds number airfoil flows, implemented in the viscous-inviscid design/analysis code ISES, is presented.
Journal ArticleDOI

Absolute and convective instabilities in free shear layers

TL;DR: The absolute or convective character of inviscid instabilities in parallel shear flows can be determined by examining the branch-point singularities of the dispersion relation for complex frequencies and wavenumbers.
Related Papers (5)
Frequently Asked Questions (21)
Q1. What have the authors contributed in "Direct numerical simulations of forced and unforced separation bubbles on an airfoil at incidence" ?

In this paper, a series of simplified linear stability analysis is presented that illustrate a three-dimensional absolute instability of the twodimensional vortex shedding that occurs naturally. 

Cusp-map method for determining the presence of absolute instabilityA simple criterion for the presence of absolute instability is the existence of an instability wave possessing zero group velocity, cg = 0, and a positive temporal growth rate, ωi > 0. 

sustained temporal disturbance growth first occurs in the region 0.5 x 0.55, suggesting that some form of absolute instability is sustained in the vicinity of the vortex shedding region. 

A domain width of at least 4 times the step height (corresponding approximately to the reattachment length) is necessary to resolve the largest spanwise structures in the case of flow over a backward-facingstep. 

The system of laminar separation, shear-layer roll-up and periodic vortex shedding gives rise to a characteristic time-averaged skin friction coefficient, cf , distribution and causes the lift coefficient, CL, to oscillate. 

Performing linear stability analysis on analytic velocity profiles similar to those observed in the DNS, Alam & Sandham found that reverse flow greater than 15 % would be required in order to sustain absolute instability, compared to an observed reverse flow of only 4–8 %. 

A series of three-dimensional simulations, resolving the linear response to three-dimensional perturbations, suggest that the two-dimensional vortex shedding behaviour is absolutely unstable to threedimensional perturbations. 

The w-perturbations grow in amplitude within individual vortices as they convect downstream; however, within the vicinity of the vortex shedding location the perturbations also exhibit growth in amplitude without convecting downstream. 

Elliptic instability is the name given to the instability of elliptical two-dimensional streamlines to three-dimensional perturbations, for which a review is given in Kerswell (2002). 

At the downstream exit boundary (ξ±), which will be subject to the passage of nonlinear fluid structures, a zonal characteristic boundary condition (Sandberg & Sandham 2006) is applied for increased effectiveness. 

The amplitude of disturbances at any fixed x-location appears to grow at the approximate rate e4t , and the amplitude of disturbances also appears to increase with increasing x-location. 

The resultant turbulent flow enhances mixing and momentum transfer in the wall-normal direction, and causes the boundary layer to reattach. 

It is important to note that upon removal of forcing, although the bubble properties change significantly, the bubble does not revert to two-dimensional behaviour. 

The first numerical simulations of separation bubbles were limited either to two-dimensional analysis (Pauley, Moin & Reynolds 1990), or only studied primary/linear instability and did not resolve transition (Pauley 1994; Rist 1994). 

The net effect is to decrease L/D from 21.1 to 17.2, hence it appears that the presence of forcing significantly improves the aerodynamic performance of the airfoil while requiring little energy input. 

Owing to the large growth rates present the probe readings were multiplied by e−σ t , whereσ = 4 is the temporal growth rate observed in the vicinity of vortex shedding, in order to better visualize the data. 

Upon removing the forcing, the turbulent behaviour can be monitored by observing pressure fluctuations within the boundary layer (figure 9). 

Thisprecludes amplification of round-off error as a route to transition, since a much larger N-factor is required to amplify round-off error (∼10−16) to nonlinear amplitudes. 

Hammond & Redekopp (1998) found that for profiles at Reδ∗ = 103, a minimum reverse flow velocity of 20 % was required to observe local absolute instability. 

The corresponding spanwise wavelengths for elliptic and mode-A instability are therefore expected to be in the range 0.15 < λ< 0.2, and the corresponding wavelength for mode-B instability is expected to be of the order λ= 0.05. 

Whilst possible for generating grids for two-dimensional simulations, an iterative grid production method is not suitable for extension to three-dimensional simulations as it would be unfeasibly expensive.