surface pressure and viscous forces on inclined elliptic

18
Surface pressure and viscous forces on inclined elliptic cylinders in steady flow SUBHANKAR SEN Department of Mechanical Engineering, Indian Institute of Technology (Indian School of Mines) Dhanbad, Dhanbad 826 004, India e-mail: [email protected] MS received 29 January 2018; revised 25 September 2019; accepted 17 March 2020 Abstract. Surface pressure characteristics of elliptic cylinders of various thicknesses and orientations are investigated in steady flow regime. A stabilized finite-element method has been used to discretize the conser- vation equations of incompressible fluid flow in two dimensions. The Reynolds number, Re, is based on the major axis of cylinder and free-stream speed. Results have been presented for Re 40 and 0 a 90 , where a is the angle of attack. Cylinder aspect ratios AR considered are 0.2 (thin), 0.5 and 0.8 (thick). It is found that a decrease in AR does not significantly alter the location of minimum surface pressure for a ¼ 90 , but the value of minimum pressure decreases sharply, resulting in severe adverse pressure gradient. In contrast, for a ¼ 0 , the location travels towards the base and the minimum pressure increases, leading to delayed flow separation. In general, the magnitude of forward stagnation pressure at low Re is smaller than the maximum pressure for AR 0:5. The maximum pressure occurs at the forward stagnation point as the Re and AR increase. However, in most cases, the locations of forward stagnation and maximum pressure points differ even when the pressure coefficients are very close to each other. The forward stagnation and maximum pressure coefficients of an elliptic cylinder decrease monotonically with increasing a. The drag of a circular cylinder in most cases exceeds the ones obtained for elliptic cylinders. With increasing AR, the drag increases approximately linearly for small a, lift decreases approximately linearly and moment decreases non-linearly. For a thick cylinder, while the effect of Re on lift and moment is insignificant, the drag shows a strong dependence. Roughly a ¼ 20 for Re ¼ 40 flow represents a critical angle of attack below which a cylinder of AR 0:5 acts like a streamlined body and above, like a bluff body. Keywords. Stabilized finite-element; elliptic cylinder; angle of attack; pressure coefficient; bluffness. 1. Introduction The flow around elliptic cylinders is associated with physics that are displayed by both streamlined as well as bluff bodies. A wide range of geometries, i.e. flat plate to circle, can be simulated by altering the ratio of length of minor to major axes or aspect ratio AR of an elliptic cylinder. The angle of attack a of an elliptic cylinder is formed between the incoming stream and its major axis (see figure 1). Symmetric configuration of an elliptic cylinder corresponds to a ¼ 0 or 90 ; intermediate values of a lead to asymmetric configura- tion. Compared with its circular (AR ¼ 1) and square coun- terparts, the analysis of flow past elliptic cylinders has received much lesser attention. Elliptic tubes find wide applications in heat exchangers owing to much lesser pumping requirements as compared with circular tubes [1]. The early investigations concerning flow past elliptic cylinders were confined mostly to semi-analytical or numerical treatment of the conservation equations at very low Reynolds number, Re. For a ¼ 0 and steady flow at Re ¼ 40, Dennis and Chang [3] predicted the drag on elliptic cylinders of varying AR. By approximately solving Oseen’s linearized equations in two dimensions, Yano and Kieda [27] presented limited results highlighting the dependence of drag coefficient C d (for AR ¼ 0:1, 0.5 and 1) and lift coefficient C l (for AR ¼ 0:1 and 0.5) on a. The results were presented for Re ¼ 0:1 and 1. With a increasing from 0 to 90 , they observed monotonic rise of C d and a non-monotonic variation of C l . By solving the Stokes equations via finite-element method, Sugihara-Seki [25] explored, for Re 0, the motion of a freely suspended elliptic cylinder in a narrow channel. Masliyah and Epstein [15] employed finite-difference method with relaxation to discretize the Navier–Stokes equations of motion cast in the streamfunction–vorticity (w x) form. For two-dimen- sional steady flow past symmetrically oriented elliptic cylinders of AR ¼ 0:2 1, they presented results for Re 90. By employing finite-difference on the w x equations, Lugt and Haussling [14] investigated the steady Sådhanå (2020)45:172 Ó Indian Academy of Sciences https://doi.org/10.1007/s12046-020-01397-z

Upload: others

Post on 16-Jun-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Surface pressure and viscous forces on inclined elliptic

Surface pressure and viscous forces on inclined elliptic cylindersin steady flow

SUBHANKAR SEN

Department of Mechanical Engineering, Indian Institute of Technology (Indian School of Mines) Dhanbad,

Dhanbad 826 004, India

e-mail: [email protected]

MS received 29 January 2018; revised 25 September 2019; accepted 17 March 2020

Abstract. Surface pressure characteristics of elliptic cylinders of various thicknesses and orientations are

investigated in steady flow regime. A stabilized finite-element method has been used to discretize the conser-

vation equations of incompressible fluid flow in two dimensions. The Reynolds number, Re, is based on the

major axis of cylinder and free-stream speed. Results have been presented for Re� 40 and 0� � a� 90�, where ais the angle of attack. Cylinder aspect ratios AR considered are 0.2 (thin), 0.5 and 0.8 (thick). It is found that a

decrease in AR does not significantly alter the location of minimum surface pressure for a ¼ 90�, but the value ofminimum pressure decreases sharply, resulting in severe adverse pressure gradient. In contrast, for a ¼ 0�, thelocation travels towards the base and the minimum pressure increases, leading to delayed flow separation. In

general, the magnitude of forward stagnation pressure at low Re is smaller than the maximum pressure for

AR� 0:5. The maximum pressure occurs at the forward stagnation point as the Re and AR increase. However, in

most cases, the locations of forward stagnation and maximum pressure points differ even when the pressure

coefficients are very close to each other. The forward stagnation and maximum pressure coefficients of an

elliptic cylinder decrease monotonically with increasing a. The drag of a circular cylinder in most cases exceeds

the ones obtained for elliptic cylinders. With increasing AR, the drag increases approximately linearly for small

a, lift decreases approximately linearly and moment decreases non-linearly. For a thick cylinder, while the effect

of Re on lift and moment is insignificant, the drag shows a strong dependence. Roughly a ¼ 20� for Re ¼ 40

flow represents a critical angle of attack below which a cylinder of AR� 0:5 acts like a streamlined body and

above, like a bluff body.

Keywords. Stabilized finite-element; elliptic cylinder; angle of attack; pressure coefficient; bluffness.

1. Introduction

The flow around elliptic cylinders is associated with physics

that are displayed by both streamlined as well as bluff bodies.

A wide range of geometries, i.e. flat plate to circle, can be

simulated by altering the ratio of length of minor to major

axes or aspect ratio AR of an elliptic cylinder. The angle of

attack a of an elliptic cylinder is formed between the

incoming stream and its major axis (see figure 1). Symmetric

configuration of an elliptic cylinder corresponds to a ¼ 0� or90�; intermediate values of a lead to asymmetric configura-

tion. Compared with its circular (AR ¼ 1) and square coun-

terparts, the analysis of flow past elliptic cylinders has

received much lesser attention. Elliptic tubes find wide

applications in heat exchangers owing to much lesser

pumping requirements as compared with circular tubes [1].

The early investigations concerning flow past elliptic

cylinders were confined mostly to semi-analytical or

numerical treatment of the conservation equations at very

low Reynolds number, Re. For a ¼ 0� and steady flow at

Re ¼ 40, Dennis and Chang [3] predicted the drag on

elliptic cylinders of varying AR. By approximately solving

Oseen’s linearized equations in two dimensions, Yano and

Kieda [27] presented limited results highlighting the

dependence of drag coefficient Cd (for AR ¼ 0:1, 0.5 and 1)and lift coefficient Cl (for AR ¼ 0:1 and 0.5) on a. Theresults were presented for Re ¼ 0:1 and 1. With aincreasing from 0� to 90�, they observed monotonic rise of

Cd and a non-monotonic variation of Cl. By solving the

Stokes equations via finite-element method, Sugihara-Seki

[25] explored, for Re � 0, the motion of a freely suspended

elliptic cylinder in a narrow channel. Masliyah and Epstein

[15] employed finite-difference method with relaxation to

discretize the Navier–Stokes equations of motion cast in the

streamfunction–vorticity (w� x) form. For two-dimen-

sional steady flow past symmetrically oriented elliptic

cylinders of AR ¼ 0:2� 1, they presented results for

Re� 90. By employing finite-difference on the w� xequations, Lugt and Haussling [14] investigated the steady

Sådhanå (2020) 45:172 � Indian Academy of Sciences

https://doi.org/10.1007/s12046-020-01397-z Sadhana(0123456789().,-volV)FT3](0123456789().,-volV)

Page 2: Surface pressure and viscous forces on inclined elliptic

and unsteady flows around thin elliptic cylinders of AR =

0.1 and 0.2 at 45� incidence. They presented detailed resultsfor fluid loading and streamline patterns at Re ¼ 15, 30 and

200. In order to explore the steady flow past a thin elliptic

cylinder of AR ¼ 0:2 at incidence, D’Alessio and Dennis

[2] used a semi-analytical method for the w� x equations.

They reported results for Re ¼ 5 and 20. Using a semi-

analytical method, Dennis and Young [4] presented, in the

steady flow regime, detailed results for the aerodynamic

coefficients of elliptic cylinders at incidence besides

exploring the separation topology. They considered thin

elliptic cylinders of AR ¼ 0:1 and 0.2. Faruquee etal [5]

numerically investigated the effects of AR on characteris-

tics of steady flow at Re ¼ 40. For elliptic cylinders of

AR ¼ 0:3� 1 and major axis parallel to the incoming

stream (a ¼ 0�), this study provides a detailed analysis of

the relationship between Cd and its pressure and viscous

components, Cdp and Cdv, respectively, with AR. For steady

unbounded flow of power-law fluids around symmetric

elliptic cylinders of AR ¼ 0:2� 1, Sivakumar etal [21]

reported extensive numerical results for fluid forces and

surface pressure besides exploring the separation phe-

nomenon. The Re range considered was 0.01-40. Sen etal

[20] presented detailed numerical results concerning steady

separation of laminar boundary layer for symmetrically as

well as asymmetrically oriented elliptic cylinders of various

AR. They proposed separation topologies for both config-

urations of the cylinder. For the inclined cylinders, signif-

icant deviation in separation topology was found relative to

those proposed earlier by Smith [22] and Dennis and Young

[4]. For elliptic cylinders of AR ¼ 0:2, 0.5 and 0.8, Sen and

Mittal [18] numerically investigated, at Re ¼ 200, the

evolution of far wake vortical structures as a function of a.Very recently, Sourav etal [24] numerically studied the

response of a freely vibrating thick elliptic cylinder of

AR ¼ 109at low Reynolds numbers.

A distinguishing feature of the surface pressure, Cp,

distribution of thin (AR of O(0.2) or smaller) elliptic

cylinders as compared with those of a circular cylinder is

that the magnitude as well as location of the maximum

pressure, Cpmax, in the former deviates from those of the

forward stagnation pressure, Cpfs, at low Re, such as 5 or

smaller. Little information is available in the literature on

this interesting phenomenon. For instance, the presence of a

maximum in Cp of an elliptic cylinder at locations other

than the leading edge or forward stagnation point was

earlier reported by Masliyah and Epstein [15] solely for the

symmetric configuration of a ¼ 90�. However, no referencewas drawn by them to the forward stagnation and maximum

pressure coefficients. At low Re, the departure of Cpmax

from Cpfs in magnitude as well as location is also evident

from figures 17b and c of Yoon etal [28] for steady flow

past a square cylinder at 0� and 45� incidences, respec-

tively. They did not report or explore this phenomenon. In

general, for steady flow around inclined elliptic cylinders of

various thicknesses, a detailed analysis of surface pressure

distribution as well as a comprehensive data set for the fluid

forces is unavailable in the literature. These research gaps

motivate the current study. In this computational work, the

features of surface pressure distribution on symmetrically

as well as asymmetrically oriented elliptic cylinders in an

unbounded medium are studied. The departure of the

maximum pressure from the forward stagnation pressure is

investigated and a possible explanation is provided. To this

end, steady laminar flow past elliptic cylinders of AR ¼ 0:2(thin), 0.5 and 0.8 (thick) at incidence have been considered

for Re� 40. For each AR, the angle of attack is varied from

0� to 90� in steps of 15�. Very limited results have also

been presented for the circular cylinder. A stabilized finite-

element solver that uses the same order of bilinear inter-

polation for fluid velocity, u ¼ ðu; vÞ, and pressure, p, has

been employed in two dimensions. The discretized lin-

earized algebraic equations have been generated using the

simultaneous or full coupling method, resulting in a single

global matrix equation system for the dependent primitive

variables, i.e. velocity and pressure.

The remaining of this article is organized in the fol-

lowing manner. In section 2, the governing differential

equations for incompressible fluid flow are reviewed. The

finite-element formulation involving SUPG (streamline-

upwind/Petrov-Galerkin) and PSPG (pressure-stabilizing/

Petrov-Galerkin) stabilization is discussed in section 3. The

definition of the problem and finite-element mesh are

depicted in sections 4 and 5, respectively. Validation of the

formulation along with mesh convergence is discussed in

section 6. The main results are presented and discussed in

section 7. In section 8, a few concluding remarks are made.

2. The governing equations

Let X � R2 represents the spatial domain and its boundary

C be piecewise smooth. The spatial coordinates are denoted

by the vector x = (x, y). In strong form, the conservation

,

,

σσ

αxx

yx

vu

u

u

v

v

x

y

= U = 0

= U

= U

= 0

= 0

= 0

= 0

u = 0 v,

baθ

80 a 120a

100a

= 0

base point

Figure 1. Problem set-up for steady unbounded flow around a

stationary elliptic cylinder inclined at an angle a to the free-

stream. The base point is marked in the figure. For this study,

AR ¼ 0:2; 0:5 and 0.8. This figure is not drawn to scale.

172 Page 2 of 18 Sådhanå (2020) 45:172

Page 3: Surface pressure and viscous forces on inclined elliptic

equations governing the steady flow of an incompressible

fluid of density, q, are

qðu � $u� fÞ � $ � r ¼ 0 on X; ð1Þ

$ � u ¼ 0 on X: ð2ÞEquations (1) and (2) are cast in vector or coordinate-free

form. The choice of stress-divergence form of the Navier–

Stokes equations (Eq. (1)) is driven by the fact that the

resultant weak form contains the surface traction vector, h(defined in Eq. (4)) as the flux or natural boundary condi-

tion term (the right hand side term of Eqs. (5) and (6)). In

these equations, f and r denote the body force per unit

volume and the Cauchy stress tensor at a point, respec-

tively. The constitutive relation for stress in terms of its

inviscid or isotropic and viscous or deviatoric contributions

reads as

r ¼ �pIþ 2leðuÞ; where eðuÞ ¼ 1

2ðð$uÞ þ ð$uÞTÞe:

ð3ÞIn Eq. (3), I, l and e stand for the identity tensor, dynamic

viscosity of the fluid and strain rate tensor, respectively.

The essential or Dirichlet and natural or Neumann-type

boundary conditions are represented as

u ¼ g on Cg; n � r ¼ h on Ch; ð4Þrespectively. Here, Cg and Ch are complementary sub-

sets of C and n is its local unit normal vector. In the

present work, the free-stream or towing tank boundary

condition is used on the lateral boundaries of the

domain (figure 1). This requires prescribed free-stream

speed condition on the upstream as well as lateral

boundaries. The no-slip condition on velocity is applied

along the surface of the cylinder. At the exit/down-

stream boundary, a Neumann condition for velocity is

specified that amounts to traction-free condition. The

boundary conditions are illustrated in figure 1. The

choice of boundary conditions as well as blockage

(defined in section 4) in the present work is represen-

tative of unbounded flow [17], [19].

3. Stabilized finite-element formulation

The spatial domain X is discretized into non-overlapping

quadrilateral elements or sub-domains Xe, e ¼ 1; 2; :::; nelwhere nel is the total number of elements. Let Sh

uand Sh

p

denote the finite-dimensional trial function spaces for u and

p, respectively. The respective weighting function spaces

are denoted by Vhuand Vh

p. The stabilized finite-element

formulation of the conservation equations (1) and (2)

combined (as in a coupled formulation) is written as

follows: find uh 2 Shuand ph 2 Sh

p such that 8wh 2 Vhu,

qh 2 VhpZ

Xwh � q uh � $uh � f

� �dX

þZXeðwhÞ : rðph; uhÞdX

þXnele¼1

ZXe

1

qsSUPGqu

h � $wh þ sPSPG$qh

� �:

q uh � $uh � f� �� $ � rðph; uhÞ� �

dXe

þZXqh$ � uhdX

þXnele¼1

ZXe

d$ � whq$ � uhdXe ¼ICh

wh � hhdC:

ð5Þ

In the combined weak or variational formulation of

Eqs. (1) and (2) presented by Eq. (5), the first, second

and fourth terms in the left hand side along with the right

hand side construct the Galerkin statement of the prob-

lem. The first series of element level integrals are the

SUPG and PSPG stabilization terms added to the varia-

tional formulations of the momentum and continuity

equations, respectively. The SUPG term provides stability

against spurious oscillations in the velocity field while

the PSPG term suppresses the spurious modes of pressure

that might appear owing to the use of the same order of

interpolation for velocity and pressure. The inclusion of

PSPG term ensures non-zero coefficient of pressure in the

discretized continuity equation and hence, introduces

definiteness in the matrix equation system. The second

series of element level integrals enhance numerical sta-

bility of the formulation at high Re. Equation (5) can be

decomposed in its respective momentum and continuity

components as

ZXwh � q uh � $uh � f

� �dX

þZXeðwhÞ : rðph; uhÞdX

þXnele¼1

ZXe

1

qsSUPGqu

h � $wh� �

:

q uh � $uh � f� �� $ � rðph; uhÞ� �

dXe

þXnele¼1

ZXe

d$ � whq$ � uhdXe ¼ICh

wh � hhdC:

ð6Þ

ZXqh$ � uhdXþ

Xnele¼1

ZXe

1

qsPSPG$q

h� �

:

q uh � $uh � f� �� $ � rðph; uhÞ� �

dXe ¼ 0:

ð7Þ

In the present work, f ¼ 0. More details of the finite-

element formulation are available in Tezduyar etal

[26].

Sådhanå (2020) 45:172 Page 3 of 18 172

Page 4: Surface pressure and viscous forces on inclined elliptic

4. The problem definition

An elliptic cylinder with major axis of length ‘a’ and minor

axis of length ‘b’ is placed in a computational domain

whose exterior boundary is chosen to be a rectangle (see

figure 1). Aspect ratio of the cylinder is computed as

AR ¼ ba. The origin of the Cartesian coordinate is located at

the centre of the fixed cylinder. The positive x axis extends

along the downstream. Relative to the negative x axis, the

angle of attack is measured clockwise whereas the cir-

cumferential angle h is measured counterclockwise. For all

a, the trailing tip represents the base point of the cylinder

(highlighted in the figure). The streamwise distance of the

upstream and downstream boundaries of the domain mea-

sured from the cylinder centre are fixed to 80a and 120a,

respectively. The lateral boundaries are equidistant from

the centre of the cylinder; the domain width or vertical

distance between these boundaries is fixed to 100a for all

the computations. This arrangement results in a blockage of

0.01 for a ¼ 90� and it is lower for a\90�. Blockage is theratio of cross-stream projection of the cylinder to the

domain width. The Reynolds number is based on the major

axis and free-stream speed, U, i.e. Re ¼ Uam where m is the

kinematic viscosity of the fluid. The moment exerted by the

fluid on the cylinder is calculated at the centre of the

cylinder. The clockwise moment is considered positive.

The following definitions are used for the fluid forces

(Cd;Cl) and moment (Cm) coefficients:

Cd ¼ 112qU2a

ZCcyl

hxdC; ð8Þ

Cl ¼ 112qU2a

ZCcyl

hydC; ð9Þ

Cm ¼ 112qU2a2

ZCcyl

r� hdC: ð10Þ

Here, hx and hy denote the streamwise and cross-stream

components, respectively, of the surface traction vector,

Ccyl represents the cylinder boundary and r is the radius

vector of any arbitrary point located on the cylinder surface

measured from the centre of cylinder.

5. Finite-element mesh

The computational domain is discretized using a non-uniform,

multiblock and structured mesh consisting of bilinear quadri-

lateral elements. Irrespective of AR and a, the number of nodes

and number of elements for each mesh are fixed to 120626 and

119768, respectively. Figure 2a shows a representative finite-

element mesh for AR ¼ 0:5 and a ¼ 45�. A close-up of the

mesh near the cylinder is shown in figure 2b. Five contiguous

mesh blocks, i.e. a central square block accommodating the

cylinder and its four neighbour rectangular blocks directed

along the left, right, top and bottom, constitute the mesh. The

central block is non-Cartesian; it is composed of two families of

orthogonal grid lines, i.e. radial and circumferential. Each cir-

cumferential grid line in the central block contains 464 nodes.

The normalized radial thickness of the first layer of elements

lying on the cylinder surface is 0.0005. The finite-element

meshesused in thiswork are identical to thoseused inour earlier

work (see Sen etal [20]).

(b)(a)

x

y

Figure 2. Steady flow past an elliptic cylinder of AR ¼ 0:5: (a) finite-element mesh corresponding to a ¼ 45� and (b) its close-up near

the cylinder. The mesh consists of 120626 nodes and 119768 bilinear quadrilateral elements. In (b), the origin of the coordinate system is

marked by a � symbol.

172 Page 4 of 18 Sådhanå (2020) 45:172

Page 5: Surface pressure and viscous forces on inclined elliptic

6. Validation and convergence of predicted results

6.1 Comparison to the earlier studies

For steady flow past elliptic cylinders, extensive validation

of fluid forces by us has been reported in Sen etal [20] and

Sourav and Sen [23].

6.2 Mesh convergence

The mesh convergence of the predicted results has been

established in one of our earlier works. Table 5 of Sen etal

[20] provides a detailed discussion on mesh convergence of

results.

7. Results

The aerodynamic behaviour of elliptic cylinders at inci-

dence is studied numerically for Re� 40 in the steady flow

regime. The aspect ratios of elliptic cylinders are 0.2, 0.5

and 0.8, while the angle of attack varies from 0� to 90�. Theelement level flow matrix and vector entries have been

computed by employing the 2� 2 points Gauss–Legendre

quadrature rule. The linearized asymmetric algebraic

equation system of flow variables has been solved by a

matrix-free GMRES or Generalized Minimal RESidual

method proposed by Saad and Schultz [16]. A diagonal pre-

conditioner has been used to accelerate the convergence of

non-linear iterations.

(a) (b)

(c)

Fig. 3. Steady unbounded flow past elliptic cylinders at Re ¼ 40: the surface vorticity distribution for cylinders of AR ¼ 0:2, 0.5 and 0.8for a ¼ (a) 0�, (b) 45� and (c) 90�. In (b), the location of forward stagnation point is highlighted for AR ¼ 0:2.

Sådhanå (2020) 45:172 Page 5 of 18 172

Page 6: Surface pressure and viscous forces on inclined elliptic

In an early numerical exploration aiming the determi-

nation of the critical Reynolds number, Rec, marking the

onset of vortex-shedding or transition to the unsteady

regime of flow, Jackson [10] considered objects of several

shapes, including circular and elliptic cylinders of varying

AR. Table 2 of Jackson [10] lists the values of Rec and

corresponding Strouhal number, Stc, for a flat plate, a cir-

cular cylinder and elliptic cylinders of various AR. From

this table, the values of Rec for AR ¼ 0:1 and 0.3 elliptic

cylinders are 29.680 and 32.765, respectively. These

numerical values underscore that the flow past the AR ¼0:2 cylinder is unsteady at the maximum Re of 40 consid-

ered in the current study. For AR ¼ 0:5, table 5 of Jackson

[10] summarizes the values of Rec and Stc with a. For

a ¼ 90�, Rec ¼ 35:704 and the values of Rec and Stcincrease monotonically as a continues to decrease to 0�.Johnson etal [11] performed two-dimensional spectral-

element computations and investigated the flow past elliptic

cylinders of AR ¼ 0:01� 1 (or major axis normal to flow)

over Re ¼ 30� 200. Figure 4 of their paper illustrating

vorticity field for AR ¼ 0:5 at Re ¼ 40 ascertains that

vortex-shedding is absent and the flow is steady. As evident

from figure 14 of Johnson etal [11], the flow for AR ¼ 0:25turns unsteady below Re ¼ 40. This figure also hints

towards a monotonic decay of Rec with reducing AR. In

view of this variation of critical Re with AR, the flow past

the AR ¼ 0:2 cylinder at Re ¼ 40 is expected to be essen-

tially unsteady. By artificially preventing the onset of vor-

tex-shedding or stabilizing the wake, a flow can, however,

remain in the steady regime at high Re (see Grove etal [8],

Fornberg, [6, 7], Sen etal [19], etc.). In the current work, we

compute the Re ¼ 40 flow past the AR ¼ 0:2 cylinder as a

steady flow. Dennis and Young [4] and Sivakumar etal [21]

also performed similar steady state computations.

7.1 Surface vorticity

Figure 3 illustrates, at Re ¼ 40, the distribution of nor-

malized surface vorticity, xDU, for symmetrically and

asymmetrically oriented elliptic cylinders of AR = 0.2, 0.5

and 0.8. It is observed that the symmetric configuration of a

cylinder leads to anti-symmetric vorticity distribution about

the base point. In this work, the aerodynamic forces have

been computed from the surface pressure and viscous

stresses along the surface. The anti-symmetry of vorticity

ensures absence of lift force for the symmetric cylinders,

i.e. a ¼ 0� and 90�. For asymmetric configurations, x dis-

plays asymmetry about the base and hence, lift force exists.

When the laminar boundary layer is attached to the cylinder

surface, the vorticity curves along the cylinder surface are

characterized by the presence of two zero-vorticity or zero-

shear stress singular points (AR = 0.2 case in figure 3a for

vorticity), i.e. the forward and rear stagnation points [13].

The appearance of two more zero-vorticity points, corre-

sponding to the separation and attachment points, implies

separated flow (AR = 0.5 and 0.8 cases in figure 3a and b,

c). The separation of laminar boundary layer from elliptic

cylinders has been discussed in Sen etal [20].

For a representative Reynolds number of 20, the classical

closed wake profiles for AR ¼ 0:2, 0.5 and 0.8 elliptic

cylinders with a ¼ 90� orientation are plotted together in

figure 4. For comparison, also shown is the wake of the

AR ¼ 1 cylinder. The largest size of wake is associated

with AR ¼ 0:2 whereas the shortest and narrowest wake

corresponds to AR ¼ 1. In particular, the length and width

of the wake decrease with increasing AR. It is discussed

later that when a ¼ 90�, the AR ¼ 0:2 cylinder is the most

bluff and the AR ¼ 0:8 cylinder is the least bluff. As shown

by Kumar etal [12], the topology of a symmetric closed

wake satisfies the kinematic constraint of Hunt etal [9]

discussed in section 7.2.

7.2 Features of surface pressure

Pressure distribution along the surface of elliptic cylin-

ders at Re = 40 is presented in figure 5 for a ¼ 0�, 45� and90�. The pressure coefficient is defined as Cp ¼ p�p0

12qU2 where

p0 is the free-stream pressure. In this work, the free-stream

pressure is assigned a zero value. Sharp spikes character-

ized by large negative value of pressure are seen in the

Cp � h curves for a ¼ 45� and 90�. At the location of

minimum pressure, Cpmin, the pressure gradient changes

sign from negative to positive or becomes adverse (also see

Faruquee etal [5]). Between the points of minimum pres-

sure, favourable pressure gradient persists along the front

surface of the cylinders and adverse pressure gradient along

the aft. The Cp � h distribution around the surface of

symmetric cylinders (figure 5a, c) is characterized by

identical pressure variation along the lower (0� � h� 180�)and upper (180� � h� 360�) surfaces. Also, in all such

cases, the forward stagnation point is located at h ¼ 0� (or

360�) and the corresponding forward stagnation pressure

represents (for Re ¼ 40) the maximum pressure on the

cylinder. The forward stagnation point is a point of zero-

Figure 4. Steady separated flow past elliptic cylinders with

major axis perpendicular to the flow in an unbounded medium:

comparison of the wake size via contours of w for AR ¼ 0:2, 0.5,0.8 and 1 at Re ¼ 20.

172 Page 6 of 18 Sådhanå (2020) 45:172

Page 7: Surface pressure and viscous forces on inclined elliptic

vorticity such that vorticity in its neighbourhood changes

sign from negative to positive (see figure 3b for AR ¼ 0:2at Re ¼ 40). Thus, location of the forward stagnation point

is determined from the surface vorticity distribution.

Details on identification of location of forward stagnation

point from surface vorticity are available in Sen etal [20].

In figures 5 and 7 of Sen etal [20], the point P denotes the

forward stagnation point. For a ¼ 0� (figure 5a), the

minimum pressure for the AR = 0.8 cylinder is obtained for

h slightly greater than 90� or smaller than 270�. However,as AR decreases, the minimum pressure points move

towards the base. Also, the value of Cpmin increases, i.e.

Cpmin becomes less negative. This is consistent with the

observations of Faruquee etal [5]. For AR ¼ 0:2, the min-

imum is achieved at the base point. The flow remains fully

attached under strong favourable pressure gradient and the

streamlines closely follow the body contour (see figure 6c).

Figure 6a shows the surface pressure and its gradient,oCp

os ,

while a close-up of the same is presented in figure 6b. The

gradient is measured along the cylinder contour (the

direction is denoted by s) between the stagnation points.

Existence of a very weak adverse pressure gradient in the

vicinity of the base point is apparent from figure 6b. When

a ¼ 90� (figure 5c), Cpmin for AR ¼ 0:2 is obtained for hslightly greater than 90� or smaller than 270�. The locationof Cpmin appears practically uninfluenced by AR (contrary

to the case of a ¼ 0�), while its value declines sharply with

decreasing AR. The severity of adverse pressure gradient

therefore increases with decreasing thickness and leads to

-2.5

-2.0

-1.0

0.0

1.0

2.0

2.5

0 90 180 270 360

Cp

θ (deg.)

AR = 0.2 0.5 0.8

-2.5

-2.0

-1.0

0.0

1.0

2.0

2.5

0 90 180 270 360

Cp

θ (deg.)

AR = 0.2 0.5 0.8

-2.5

-2.0

-1.0

0.0

1.0

2.0

2.5

0 90 180 270 360

Cp

θ (deg.)

AR = 0.2 0.5 0.8

(a) (b)

(c)

90

45a = 0

Figure 5. Steady unbounded flow at Re ¼ 40 past elliptic cylinders of AR ¼ 0:2, 0.5 and 0.8: distribution of surface pressure for a = (a)

0�, (b) 45� and (c) 90�.

Sådhanå (2020) 45:172 Page 7 of 18 172

Page 8: Surface pressure and viscous forces on inclined elliptic

earlier (at lower Re) separation of the laminar boundary

layer. For AR ¼ 0:2, figure 7 shows strong adverse pressure

gradient near the cylinder shoulders (h � 90� or 270�) andthe wake comprising an attached separation bubble.

Asymmetry about the h ¼ 180� location in the Cp � hcurves is introduced when the cylinders are oriented

asymmetrically (figure 5b). For each AR the minimum

pressure on the upper rear surface is smaller than the

minimum pressure on the lower front surface, implying

stronger adverse pressure gradient on the upper rear sur-

face. This effect is more pronounced with reducing thick-

ness. The magnitude of minimum pressure of a symmetric

elliptic cylinder decreases with increasing Re (see figure 8a

and c for AR ¼ 0:2). For the range of Re considered, it is

also noted that the location of Cpmin for symmetric con-

figurations of the AR ¼ 0:2 cylinder, i.e. a ¼ 0� and 90�, ispractically insensitive to Re. For a ¼ 0�, the minimum is

attained at the base point (figure 8a) and for a ¼ 90�, at theshoulders (figure 8c and seventh column of table 1). As

evident from figure 8b for a ¼ 45�, the location of Cpmin for

asymmetric orientations, in contrast, depends on Re. The

location of Cpmin for symmetric cylinders of AR ¼ 0:5 and

0.8 moves upstream with increasing Re (not shown).As

stated in section 7.1, Sen etal [20] studied the steady sep-

aration of laminar boundary layer from inclined elliptic

cylinders of AR ¼ 0:2, 0.5 and 0.8. They explored the

separation topology for symmetric as well as asymmetric

configurations of the cylinders. As outlined in Kumar etal

[12], the wake structure for symmetric configurations

relates to the classical closed wake comprising a pair of

counter-rotating eddies (e.g. figures 4, 6c, 7c and 16d of

the present paper) whereas for asymmetric orientations, the

wake is open and consists of an attached (to the rear of the

cylinder) and a detached vortex (figure 9a of the present

paper for AR ¼ 0:8; a ¼ 45� and Re ¼ 30). The open wake

for asymmetric cylinders represents an instance of alleyway

flow. Streamline patterns for elliptic cylinders with sym-

metric/asymmetric configurations were presented in fig-

ures 3 and 4 of Sen etal [20]; schematics of the separation

topology for both types of orientations were shown in

-30

-25

-20

-15

-10

-5

0

5

0 90 180 270 360

Cp,

dCp/

ds

θ (deg.)

CpdCp/ds

-0.8

-0.6

-0.4

-0.2

0

0.2

0 90 180 270 360

s

s

(b)(a) (c)

U

Figure 6. Steady unbounded flow past an elliptic cylinder of AR ¼ 0:2 for a ¼ 0� at Re ¼ 40: (a) distribution of surface pressure andoCp

os along the cylinder contour, (b) close-up of (a) and (c) streamlines indicating attached flow.

(b)(a) (c)

-35-30-25-20-15-10-5 0 5

10 15

0 90 180 270 360

Cp,

dCp/

ds

θ (deg.)

CpdCp/ds

-2

-1

0

1

2

0 90 180 270 360

Figure 7. Steady unbounded flow past an elliptic cylinder of AR ¼ 0:2 for a ¼ 90� at Re ¼ 40: (a) distribution of surface pressure and

its gradient along the cylinder contour, (b) close-up of (a) and (c) streamlines indicating separated flow with a closed attached wake.

172 Page 8 of 18 Sådhanå (2020) 45:172

Page 9: Surface pressure and viscous forces on inclined elliptic

figure 5. The cylinder contour, known as the surface

streamline, is conventionally treated as the zero (reference)

streamline. Thus, the value of the streamfunction is zero for

this streamline. All streamlines emanating from or

terminating to the reference streamline are zero streamlines

and the points of intersection of streamlines are known as

critical points, where the vorticity and shear stress disap-

pear. The critical points are broadly classified as nodes and

-3

0

3

6

0 90 180 270 360

Cp

θ (deg.)

Re = 52040

-4

-3

-2

-1

0

1

2

3

4

5

0 90 180 270 360

Cp

θ (deg.)

-5

-4

-3

-2

-1

0

1

2

0 90 180 270 360

Cp

θ (deg.)

(a)

(c)

(b)

45

90

0α =

Figure 8. Distribution of surface pressure for steady unbounded flow past an elliptic cylinder of AR ¼ 0:2 for Re ¼ 5; 20 and 40 and

a ¼ (a) 0�, (b) 45� and (c) 90�.

Table 1. Steady unbounded flow past an elliptic cylinder of AR ¼ 0:2; a ¼ 90� for Re ¼ 1; 6; 7 and 40: summary of the values of

maximum and minimum Cp along with the h location. Also listed in the last column are the values of Cp at the shoulder. For Re� 6,

Cpfs\Cpmax and for Re 7, Cpfs ¼ Cpmax.

AR Re Cp0 Cpmax h for Cpmax Cpmin h for Cpmin Cp at h ¼ 90�

0.2 1 3.7115 7.5124 86.90� –9.9810 92.33� –2.8959

0.2 6 1.5624 1.6256 86.12� –4.1923 91.55� –2.5845

0.2 7 1.4900 1.4900 0� –3.9376 91.55� –2.5489

0.2 40 1.1024 1.1024 0� –1.8616 90.78� –1.7155

Sådhanå (2020) 45:172 Page 9 of 18 172

Page 10: Surface pressure and viscous forces on inclined elliptic

saddles. The zero streamlines along with the critical or

singular points constitute the separation topology. Regard-

ing kinematic stability of a separation topology, Hunt etal

[9] presented the following constraint to be satisfied:

XN þ 1

2

XN 0

� ��

XSþ 1

2

XS0

� �¼ 1� m: ð11Þ

The quantities N and S appearing in Eq. (11) signify the

four-way nodes and four-way saddles, respectively. The

three-way nodes and three-way saddles are, respectively,

denoted by symbols N 0 and S0. The connectivity of the flow

section is denoted by m; in the context of flow past a

solitary obstacle, m ¼ 2. For the AR ¼ 0:8 cylinder with

a ¼ 45�, figure 9a depicts, via streamlines, the alleyway

flow at Re ¼ 30. The corresponding separation topology is

shown schematically in figure 9b, where A and S, respec-

tively, stand for attachment and separation points that

appear alternately along the fluid–solid interface. They are

no-slip type three-way (or half) saddle points. The eddy

centres, denoted by N1 and N2, are four-way nodes. The

sole four-way saddle point, denoted by M, exists in the fluid

medium. Point M is the asymmetric equivalent to the wake

stagnation point that exists in a symmetric, closed wake

(marks the closure of the wake). With this identification of

critical points,P

S ¼ 1 (point M),P

N ¼ 2 (points N1 and

N2),P

N 0 ¼ 0 andP

S0 ¼ 4 (points A1, A2, S1 and S2).

The left hand side of Eq. (11) reduces to

2þ 0� 1� 42¼ �1. The right hand side, i.e. 1� m, also

yields the same value of �1. Thus, the separation topology

of alleyway flow proposed by Sen etal [20] is stable.

For Re ¼ 5� 40, figure 10 depicts the relationship

between the forward stagnation pressure coefficient, Cp0,

and Re. Irrespective of AR and a, the pressure coefficient

drops with Re. For a ¼ 90�, Cp0 appears to be the least

sensitive to AR. The relationship between base suction,

�Cpb, and Re is illustrated in figure 11a, b and c for

a ¼ 0�, 45� and 90�, respectively. For fixed AR and a, thebase suction of a cylinder decreases with increasing Re. In

general, �Cpb for a fixed Re decreases with increasing AR.

The decay of base suction signifies recovery of pressure at

the base of the cylinder. This eventually leads to lower

pressure drag.

Figures 12a, b and 13a–c, respectively, demonstrate that

Cpmax for low Re does not occur at the forward stagnation

point of symmetric (here, a ¼ 90�) and asymmetric cylin-

ders of AR� 0:5. As revealed by figure 12a, the maximum

surface pressure on a thin cylinder with a ¼ 90� at low Re,

i.e. Re� 6, occurs near its shoulders instead of the forward

stagnation point. As Re is increased, the maximum pressure

point travels upstream along the cylinder surface (or w ¼ 0

streamline) and coincides with the forward stagnation point

for Re[ 6. The departure of Cpfs from Cpmax is also dis-

played by the AR ¼ 0:5 cylinder at Re ¼ 1 (see the inset in

figure 12b), where the maximum pressure points are loca-

ted upstream of the shoulders. With increasing cylinder

thickness (AR ¼ 0:8, for instance), the difference between

Cpfs and Cpmax diminishes (inset of figure 12c) and the

maximum pressure point merges with the forward stagna-

tion point for all Re. As mentioned in section 1, the pres-

ence of a maximum in Cp at low Re was earlier reported by

Masliyah and Epstein [15] for AR ¼ 0:2 and a ¼ 90�.Though it is not explicitly pointed out, similar observations

can be made from the pressure distribution curves presented

by Sivakumar etal [21] for AR = 0.2 and a ¼ 90� at very

low Re (= 0.01). For Re ¼ 5 and 40, figure 13a–c demon-

strates that both Cpfs and Cpmax decrease monotonically

with increasing a. The (Cpmax � Cpfs) difference decreases

with increasing Re as well as AR, and approaches zero for

the thick cylinder. It is also noted that both Cpfs and Cpmax

decrease with increasing Re. The dependence of h locations

of the forward stagnation (hfs) and maximum (hmax) pres-sure points on a is depicted via figure 13d–f. For

0� � a� 90�, both hfs and hmax, in general, display non-

monotonic variation with a. A departure from the non-

monotonic hmax � a trend is found to occur for a thin

cylinder at Re ¼ 5 (figure 13d), which shows linear

decrease of hmax with a. Interestingly, hfs and hmax differ inmost cases (such that hfs [ hmax or the maximum pressure

point resides above the forward stagnation point) even

when the values of Cpfs and Cpmax are very close to each

MN

NA1

S1

A

S

1

22

2

(a) (b)

ψ = 0

Figure 9. Steady separated flow past an inclined elliptic cylinder of AR ¼ 0:8 with a ¼ 45�: (a) streamlines depicting the asymmetric

wake at Re ¼ 30 and (b) schematic of corresponding separation topology illustrating the critical points.

172 Page 10 of 18 Sådhanå (2020) 45:172

Page 11: Surface pressure and viscous forces on inclined elliptic

other. For an asymmetric cylinder, the second row of fig-

ure 13 indicates that the difference between hfs and hmaxdecreases with increasing Re as well as AR.

This mismatch in magnitude as well as location of for-

ward stagnation pressure and maximum pressure at low Re

appears to be governed by a pair of factors, i.e. the Rey-

nolds number and the streamwise extent of the body. A low

value of Re and abrupt change of body contour appear to

facilitate the occurrence of this phenomenon. Following the

discussion in section 1 it is obvious that the shoulders of

diamond, thin elliptic cylinders and front corners of a

square cylinder are ideal locations where Cp at low Re turns

the maximum and hence, exceeds Cpfs. It may be noted that

the mismatch is absent for a ¼ 0� even at an Re as small as

0.01 (see the first row of figure 12 of Sivakumar etal [21]

corresponding to the value of the power-law index being

unity). For this configuration, the streamwise extent of the

cylinder is appreciably high. At low Re, the mode of

diffusion transport dominates and the pressure scales with

U. As Re is increased, transport of vorticity by convection

becomes appreciable. In addition, the order of magnitude of

pressure equals the square of U. For AR ¼ 0:2 and 0.8

cylinders at zero incidence, figure 14a and b, respectively,

depicts the relationship of logarithm of Cp0 and Cpb with

logarithm of Re. Since the pressure coefficients decay with

Re or follow an inverse relationship with Re (figures 10

and 11), the log Cp–log Re variation in figure 14 is almost

linear with a negative slope. Thus, for the range of Re

considered, the flow is in the viscous regime and pressure

varies with U. For flow past a symmetric bluff body, the

adverse pressure gradient generally commences at locations

close to its shoulders (see figure 7a and b for instance).

Thus, near the shoulders, the magnitude of velocity is low.

For AR ¼ 0:2, the flow around a shoulder becomes equiv-

alent to external flow past a corner. The cylinder surface is

a zero streamline, along which u ¼ 0 by virtue of the no-

1

2

3

4

5

6

0 10 20 30 40

Cp0

Re

AR = 0.2 0.5 0.8

0.5

1.0

1.5

2.0

0 10 20 30 401.0

1.2

1.4

1.6

1.8

2.0

0 10 20 30 40

a =

)c()b()a(

0 45 90

Figure 10. Steady unbounded flow past elliptic cylinders of AR ¼ 0:2, 0.5 and 0.8 for Re ¼ 5� 40: variation of the forward stagnation

pressure coefficient with Re for a ¼ (a) 0�, (b) 45� and (c) 90�.

0

1

2

3

0 10 20 30 40

-Cpb

Re

AR = 0.2 0.5 0.8

0.5

1.5

2.5

3.5

0 10 20 30 40 0.4

0.8

1.2

0 10 20 30 40

(a) (c)(b)

α = 0

0954

Figure 11. Steady unbounded flow past elliptic cylinders of AR ¼ 0:2, 0.5 and 0.8 for Re ¼ 5� 40: variation of the base suction

coefficient with Re for a ¼ (a) 0�, (b) 45� and (c) 90�.

Sådhanå (2020) 45:172 Page 11 of 18 172

Page 12: Surface pressure and viscous forces on inclined elliptic

-11

-4

0

4

8

0 90 180 270 360

Cp

θ (deg.)

Re = 1 3 6 7

-5

-4

-2

0

2

4

0 90 180 270 360

Re = 1 2

3.87

3.89

3.91

0 20 40 60

-4

-2

0

2

4

5

0 90 180 270 360

Re = 1

4.06

4.07

4.08

0 4 8

1.3

1.5

1.7

85 86 87 88

)c()b()a(

AR = 8.05.02.0

Figure 12. Steady unbounded flow past elliptic cylinders for a ¼ 90�: the surface pressure distribution at low Re for AR ¼ (a) 0.2, (b)

0.5 and (c) 0.8. For AR ¼ 0:2 at Re ¼ 7, it should be noted that the value of Cp at the local peak around h ¼ 86� (inset of figure a) is quitesmaller than the corresponding value of Cpfs summarized in table 1.

1

2

3

4

5

6

0 30 60 90

Cpf

s,C

pmax

α (deg.)

Re = 5, Cpfs Cpmax

Re = 40, Cpfs Cpmax

1.0

1.5

2.0

2.5

3.0

0 30 60 90

270

300

330

360

0 30 60 90

θ fs,

θ max

(de

g.)

α (deg.)

Re = 5, θfs θmax

Re = 40, θfs θmax

320

330

340

350

360

0 30 60 90

1.0

1.3

1.6

1.9

2.2

0 30 60 90

348

352

356

360

0 30 60 90

)c()b()a(

(d) (e) (f)

AR = 0.2

AR = 0.20.5

0.5 0.8

0.8

Figure 13. Steady unbounded flow past elliptic cylinders at Re ¼ 5 and 40: variation of the forward stagnation and maximum pressure

with a for AR ¼ (a) 0.2, (b) 0.5 and (c) 0.8. Also shown is the variation of h location of the forward stagnation and maximum pressure

points with a for AR ¼ (d) 0.2, (e) 0.5 and (f) 0.8.

172 Page 12 of 18 Sådhanå (2020) 45:172

Page 13: Surface pressure and viscous forces on inclined elliptic

slip condition. Thus, application of Bernoulli’s theorem

(though not applicable for real fluids, Bernoulli’s principle

provides useful estimates of velocity and pressure in real

flow) along this surface streamline cannot explain the dis-

crepancies between Cpfs and Cpmax. The streamlines next to

the w ¼ 0 streamline should therefore be considered to this

end. The streamlines upstream and downstream of the

corner, i.e. cylinder shoulder, undergo significant diversion

or bending. This excessive abrupt bending over an extre-

mely short extent (at the shoulder) is associated with a local

regime of low velocity around (ahead of and behind) the

shoulder. Thus, regimes of pressure of large magnitude, in

accordance with the principle of Bernoulli, exist in front of

and behind the shoulder. The pressure along the front half

of the cylinder is positive, and a negative or suction pres-

sure prevails along its rear half. For Re ¼ 1, the pressure at

the shoulders is negative and as Re increases, the h extent ofnegative pressure travels upstream. Large positive pressure

and large negative pressure, respectively, ahead and behind

the shoulder (see table 1) result in a discontinuity in Cp, i.e.

a positive and a negative pressure spike appear around

h ¼ 90�. The formation of the positive pressure spike is

responsible for the shift of the maximum pressure near the

shoulder rather than at the forward stagnation point. The

diversion of streamlines is most dominant at low Re regime

of attached flow, and both spikes are sharp. As Re increa-

ses, the average velocity of flow around the cylinder also

increases. Besides, a wake bubble first appears, then

enlarges and displaces its adjacent streamlines laterally

outwards. This effect reduces the bending of neighbour

streamlines. Because of the formation of wake, the average

velocity of flow behind the cylinder is smaller than that

ahead of the cylinder. The positive pressure spike is thus

eliminated or smoothed out, rendering the maximum pres-

sure to be attained at the forward stagnation point. This

leads to the identical values of Cpfs and Cpmax at higher Re.

For AR ¼ 0:5 and 0.8, the variation of velocity from near

the leading edge to the shoulders occurs in a gradual

manner. This appears to result in much lesser divergence in

h location of forward stagnation and maximum pressures

for these cylinders.

In summary, the angle by which the upstream

streamlines close to the cylinder shoulder get deflected

appears to govern whether Cpfs and Cpmax are identical or

different. For flow around the AR ¼ 0:2 cylinder at low

Re, the angle of diversion is very small (\\90�) and

Cpfs\Cpmax. As Re increases, the wake bubble expands

and the angle of diversion exceeds 90�. Consequently,Cpfs ¼ Cpmax. In case of a circular cylinder, the angle of

diversion is obtuse regardless of the flow being attached

or separated. In this case, Cpfs essentially represents

Cpmax for all Re.

7.3 Fluid loading

The dependence of drag coefficient on Reynolds number is

illustrated in figures 15a–c for cylinders of AR ¼ 0:2, 0.5,0.8 and 1. Three different values of a, i.e. 0�, 45� and 90�,are considered. Regardless of aspect ratio and angle of

attack, a monotonic decrease of Cd with Re is observed. For

a ¼ 0� and 45�, the Cd � Re curves for different cylinder

shapes are almost parallel and when Re is held constant, Cd

exhibits monotonic decrease with decreasing AR. However,

a contrasting feature is observed for a ¼ 90� (figure 15c)

where the effect of cylinder shape on Cd appears relatively

less significant. Figure 15d, e and f for, respectively, Re =

5, 20 and 40 indicates that Cd increases approximately

linearly with AR for small values of a. However, for large a,a slow decrease of Cd with increasing AR is apparent for

Re 20. The drag coefficient as a function of a is plotted in

figure 15g, h and i for Re ¼ 5, 20 and 40, respectively. For

each Re, drag of a cylinder increases monotonically with afrom its minimum at a ¼ 0� to the maximum at a ¼ 90�.This trend is consistent with the observations of Yano and

0.1

1

10

1 10 100

Cp0

, Cpb

Re

Cp0Cpb

0.1

1

10

1 10 100

Cp0

, Cpb

Re

Cp0Cpb

0.2 0.8

(b)(a)

AR =

Figure 14. Variation of Cp0 and Cpb with Re in log–log plane for steady unbounded flow past elliptic cylinders of AR ¼ (a) 0.2 and (b)

0.8 both at 0� incidence.

Sådhanå (2020) 45:172 Page 13 of 18 172

Page 14: Surface pressure and viscous forces on inclined elliptic

Kieda [27] for Re� 1 and Dennis and Young [4] for a thin

cylinder at Re ¼ 5, 20 and 40. At low Re, such as 5, the

drag of a circular cylinder overshadows those of its elliptic

counterparts of any orientation. Also, over a considerable

fraction of the range of a, the drag on an elliptic cylinder

for Re 20 (figure 15h, i) is less than the one for a circular

cylinder. The angle of attack at which elliptic cylinders of

various AR generate approximately the same drag has an

inverse relationship with Re. For instance, at Re ¼ 10,

cylinders of various thicknesses give the same drag when

a ¼ 90� (figure 15c). For Re ¼ 20 and 40, these angles

approximately equal 75� (figure 15h) and 60� (figure 15i),

respectively.

For a streamlined body, Cdv is dominant over Cdp. The

value ofCdp

Cdvratio for such bodies, therefore, falls below

unity. In contrast, for a bluff body, the value ofCdp

Cdvratio

exceeds unity. For Re ¼ 5, 20 and 40, figure 16a, b and c,

respectively, plots the relationship between theCdp

Cdvratio and

a. For each AR and Re, theCdp

Cdvratio increases monotonically

as a continues to rise. For Re ¼ 5, the value ofCdp

Cdvratio for

each AR is virtually constant at unity for a � 34�. For a

smaller than this angle,Cdp

Cdv\1, signifying that the stream-

lined nature dominates; for higher values of a, the cylinders

tend to become more bluff. When Re ¼ 20, theCdp

Cdvratio

turns invariant to AR for a � 30�. For elliptic cylinders withAR� 0:5, the streamlined nature becomes predominant for

values of a smaller than 26�, approximately. For Re ¼ 40,

the angle at which the ratio of pressure to viscous drag

0.5

1.5

2.5

3.5

4.5

0 10 20 30 40

Cd

Re

AR = 0.2 0.5 0.8

1

0.5

1.5

2.5

3.5

4.5

0 10 20 30 40 0.5

1.5

2.5

3.5

4.5

0 10 20 30 40

2.4

3.0

3.6

4.2

0.2 0.4 0.6 0.8 1

Cd

AR

α = 0ο

15ο

45ο

75ο

90ο

1.0

1.4

1.8

2.2

0.2 0.4 0.6 0.8 10.6

0.9

1.2

1.5

1.8

0.2 0.4 0.6 0.8 1

2.6

3.0

3.4

3.8

4.2

0 30 60 90

Cd

α (deg.)

AR = 0.2 0.5 0.8

1

1.0

1.4

1.8

2.2

0 30 60 900.7

1.0

1.3

1.5

1.7

0 30 60 90

(a)

)h()g(

)f()e()d(

)c()b(

α = 0

Re = 50402

0954

20 40Re = 5

(i)

Figure 15. Steady unbounded flow past elliptic cylinders of aspect ratios 0.2, 0.5, 0.8 and 1: variation of total drag with Re for a = (a)

0�, (b) 45� and (c) 90�; with AR for Re ¼ (d) 5, (e) 20 and (f) 40; with a for Re ¼ (g) 5, (h) 20 and (i) 40.

172 Page 14 of 18 Sådhanå (2020) 45:172

Page 15: Surface pressure and viscous forces on inclined elliptic

components displays insensitivity to Re lowers down to

about 30�. It may be noted that the value ofCdp

Cdvratio at this

angle exceeds unity. With AR decreasing, the ratio at a

given angle of attack decreases for a� 20� and increases

for a 45�. Roughly the neighbourhood of 20� represents acritical zone of angle of attack, below which a cylinder (for

AR� 0:5) tends to behave like a streamlined body and

above, like a bluff body. Therefore, the regime of a over

which an elliptic cylinder dominantly exhibits the features

of a bluff body rather than a streamlined body increases

with Re. However, for Re ¼ 5; 20 and 40 each, the con-

figurations of a ¼ 90� and 0�, respectively, correspond to

the maximum and minimum bluffness of an elliptic cylin-

der. At the terminal values of a, i.e. a ¼ 90� and 0�,respectively, the maximum bluff and maximum streamlined

behaviours are associated with the AR ¼ 0:2 elliptic

cylinder. An opposite trend is displayed by the thick

cylinder of AR ¼ 0:8. For a ¼ 90� configuration, figure 16d

shows the wake of the AR ¼ 0:2 cylinder at a very low Re

value of 1.07 further underscoring the fact that the maxi-

mum bluff behaviour is associated with this cylinder.

A cylinder placed symmetrically, relative to the free-

stream, experiences no net lift or moment. This feature is an

outcome of symmetric surface pressure (figure 5a, c) and

anti-symmetric surface vorticity (figure 3a, c) distributions

about the base point. Figure 17a, b and c illustrates the

dependence of Cl on Re for a = 15�, 45� and 75�, respec-tively. A monotonically decreasing Cl � Re relationship

that depends strongly on the cylinder shape is observed. For

a thick cylinder (AR = 0.8), the decay of Cl with Re is

insignificant. In contrast, Cd shows a strong decrease.

Figure 17d–f demonstrates, for moderate to higher values

of a, approximate linear decrease of Cl with increasing AR.

The Cl � a variation for Re ¼ 5, 20 and 40 is shown,

respectively, in figure 17g, h and i for AR = 0.2, 0.5 and 0.8.

As a increases from 0� to 90�, Cl exhibits non-monotonic

variation with a, i.e. initially increases from and then

decreases to zero. Similar variation of Cl with a was earlier

(a) (b)

(c) (d)

Figure 16. Steady unbounded flow past elliptic cylinders of AR ¼ 0.2, 0.5 and 0.8: variation of the Cdp=Cdv ratio with a for Re ¼ (a) 5,

(b) 20 and (c) 40. The approximate boundary of a, roughly demarcating the regimes of streamlined and bluff behaviours, is also shown. It

may be noted that this boundary shifts towards lower a as Re is increased progressively. In terms of streamlines, figure d shows the closed

wake of the AR ¼ 0:2 cylinder for a ¼ 90� at Re ¼ 1:07.

Sådhanå (2020) 45:172 Page 15 of 18 172

Page 16: Surface pressure and viscous forces on inclined elliptic

noted by Yano and Kieda [27] and Dennis and Young [4].

Irrespective of Re, the maxima in Cl for AR 0:5 occur at

a ¼ 45� among the angles of attack studied. For the thin

cylinder, however, the maximum is attained at a ¼ 45� forRe ¼ 5 and at a approximately 30� for Re ¼ 20; 40. It isalso observed that the Cl � a curves for AR ¼ 0:8 are

nearly symmetric about a ¼ 45�. With decreasing AR, lack

of symmetry about a ¼ 45� becomes prominent; Cl pre-

dicted at a\45� overshadows the corresponding value at

90� � a, i.e. the maximum lift shifts to \45�.For Re ¼ 40, table 2 summarizes the values of aerody-

namic coefficients along with their pressure and viscous

components. For symmetric elliptic cylinders, the mean lift

is zero. The viscous components of lift cancel out for such

orientations (opposite signed between the upper and lower

surfaces). For other orientations, the surface pressure

between the stagnation points displays asymmetry; hence,

lift is non-zero. For such cases, negative sign is associated

with the viscous component of lift force, Clv. The negative

or downward acting Clv was earlier predicted by Lugt and

Haussling [14] and Dennis and Young [4]. It is found that

Clp is an order of magnitude higher than Clv.

8. Conclusion

A stabilized finite-element method has been employed to

investigate the aerodynamic loading for steady, laminar

flow around stationary elliptic cylinders at various angles of

attack. The aspect ratios studied are 0.2, 0.5 and 0.8. The

aerodynamic behaviour is studied for Re� 40 and

0� � a� 90�. With major axis of the cylinder normal to the

flow, the forward stagnation pressure for AR ¼ 0:2, Re ¼1� 6 and AR ¼ 0:5, Re ¼ 1 is always smaller than the

maximum pressure occurring at/near the cylinder shoulders.

The discrepancy between Cpfs and Cpmax exists for low Re

flow around cylinders having abrupt changes in body con-

tour. With increasing Re and AR, the forward stagnation

pressure represents the maximum pressure for both the

symmetric (here, a ¼ 90�) and asymmetric configurations.

The angle of diversion of the upstream streamlines close to

the cylinder shoulder appears to govern the occurrence of

this phenomenon. For Re on the order of unity, diversion by

small angles renders Cpfs to fall short of Cpmax; these

quantities become identical when the angle of diversion is

obtuse, as in a circular cylinder. The closeness of forward

stagnation and maximum pressure, however, does not

imply that they occur at the same location on the surface of

an asymmetric cylinder. The forward stagnation and max-

imum surface pressure of a cylinder continue to decrease

with increasing a. For symmetric cylinders, the magnitude

of minimum surface pressure decreases with increasing Re.

While the location of minimum pressure, or alternatively

onset of adverse pressure gradient, travels upstream for

AR ¼ 0:5 and 0.8, the location is fixed at the base point (for

a ¼ 0�) or shoulders (for a ¼ 90�) for AR ¼ 0:2. When Re

and a are both fixed, the base suction coefficient generally

decreases with increasing AR. With increasing Re, the base

Table 2. Steady unbounded flow past elliptic cylinders of 0:2�AR� 0:8: summary of the aerodynamic coefficients and their com-

ponents at Re ¼ 40 for various angles of attack between 0� and 90�. It may be noted that the moment disappears when the lift is zero.

AR a Cdp Cdv Cd Clp Clv Cl Cmp Cmv Cm

0.2 0� 0.1648 0.6105 0.7753 0 0 0 0 0 0

15� 0.3106 0.5381 0.8487 0.6393 –0.0525 0.5868 0.1664 0.0132 0.1796

30� 0.6329 0.4043 1.0372 0.9055 –0.0748 0.8306 0.2372 0.0151 0.2523

45� 0.9662 0.2969 1.2631 0.8567 –0.0810 0.7757 0.2205 0.0109 0.2314

60� 1.2394 0.2204 1.4598 0.6480 –0.0702 0.5778 0.1613 0.0066 0.1679

75� 1.4299 0.1784 1.6083 0.3512 –0.0516 0.2996 0.0826 0.0049 0.0875

90� 1.4792 0.1559 1.6351 0 0 0 0 0 0

0.5 0� 0.4488 0.5840 1.0328 0 0 0 0 0 0

15� 0.5154 0.5572 1.0726 0.3211 –0.0179 0.3032 0.1121 0.0099 0.1220

30� 0.6842 0.4945 1.1787 0.5038 –0.0297 0.4741 0.1769 0.0139 0.1908

45� 0.8901 0.4264 1.3165 0.5196 –0.0344 0.4852 0.1820 0.0123 0.1943

60� 1.0749 0.3708 1.4457 0.4111 –0.0308 0.3803 0.1423 0.0084 0.1507

75� 1.2001 0.3352 1.5353 0.2247 –0.0183 0.2064 0.0768 0.0041 0.0809

90� 1.2443 0.3228 1.5671 0 0 0 0 0 0

0.8 0� 0.7669 0.5473 1.3142 0 0 0 0 0 0

15� 0.7891 0.5398 1.3289 0.1081 –0.0052 0.1029 0.0476 0.0035 0.0511

30� 0.8487 0.5202 1.3689 0.1814 –0.0091 0.1723 0.0798 0.0057 0.0855

45� 0.9274 0.4951 1.4225 0.2014 –0.0107 0.1907 0.0885 0.0060 0.0945

60� 1.0032 0.4719 1.4751 0.1682 –0.0095 0.1587 0.0737 0.0047 0.0784

75� 1.0570 0.4558 1.5128 0.0947 –0.0056 0.0891 0.0414 0.0026 0.0440

90� 1.0764 0.4501 1.5265 0 0 0 0 0 0

172 Page 16 of 18 Sådhanå (2020) 45:172

Page 17: Surface pressure and viscous forces on inclined elliptic

suction continues to decay. As a increases from 0� to 90�,irrespective of AR, the drag coefficient increases mono-

tonically while lift and moment increase from and then

decrease to zero. As AR increases to unity, the lift and

moment decrease monotonically to zero. In contrast, the

drag increases or decreases slowly depending on Re and a.The drag of a circular cylinder at low Re is more than that

generated by an elliptic cylinder. As Re increases, this trend

prevails over a wide range of a. The regime of a over whichan elliptic cylinder displays the features of a bluff body,

more dominantly than those of a streamlined body,

increases with Re. For Re ¼ 40, elliptic cylinders of

AR� 0:5 behave more like a streamlined body for a\20�

and a bluff body for a 20�.

Nomenclaturea Major axis of ellipse

b Minor axis of ellipse

AR Aspect ratio of ellipse

Cd Drag coefficient

Cdp Pressure drag coefficient

Cdv Viscous drag coefficient

Cl Lift coefficient

Clp Pressure lift coefficient

Clv Viscous lift coefficient

Cm Moment coefficient

Cmp Pressure component of moment coefficient

Cmv Viscous component of moment coefficient

Cp Pressure coefficient

0.0

0.4

0.8

1.2

0 10 20 30 40

Cl

Re

AR = 0.2 0.5 0.8

0.0

0.4

0.8

1.2

0 10 20 30 400.0

0.4

0.8

1.2

0 10 20 30 40

0.0

0.4

0.8

1.2

0 0.2 0.4 0.6 0.8 1

Cl

AR

α = 15ο

30ο

45ο

60ο

75ο

0.0

0.4

0.8

1.2

0 0.2 0.4 0.6 0.8 10.0

0.4

0.8

1.2

0 0.2 0.4 0.6 0.8 1

0.0

0.4

0.8

1.2

0 30 60 90

Cl

α (deg.)

AR = 0.2 0.5 0.8

0.0

0.4

0.8

1.2

0 30 60 900.0

0.4

0.8

1.2

0 30 60 90

(a) (c)(b)

5754

(d) (f)(e)

Re = 0402 Re = Re = 5

Re = 20

(g)

= α = αα = 15

5Re = Re = 40

(i)(h)

Figure 17. Steady unbounded flow past elliptic cylinders of AR ¼ 0:2, 0.5, 0.8 and 1: variation of total lift with Re for a ¼ (a) 15�, (b)45� and (c) 75�; with AR for Re ¼ (d) 5, (e) 20 and (f) 40; with a for Re ¼ (g) 5, (h) 20 and (i) 40.

Sådhanå (2020) 45:172 Page 17 of 18 172

Page 18: Surface pressure and viscous forces on inclined elliptic

Cpb Base pressure coefficient

Cpfs Forward stagnation pressure coefficient

Cpmax Maximum value of surface pressure coefficient

Cpmin Minimum value of surface pressure coefficient

h Surface traction vector

m Connectivity of flow section

n Unit normal to the surface

N Number of four-way nodes

N0

Number of three-way nodes

p Pressure

q Weight function for pressure

Re Reynolds number

S Number of four-way saddles

S0

Number of three-way saddles

u Velocity vector

U Free-stream speed

x Spatial coordinate vector

w Weight function vector for velocity components

a Angle of attack

d Stabilization parameter

� Strain rate tensor

q Density of fluid

w Streamfunction

x Vorticity

h Circumferential angle

r Stress tensor

X Domain

C Domain boundary

References

[1] Castiglia D, Balabani S, Papadakis G and Yianneskis M

2001 An experimental and numerical study of the flow past

elliptic cylinder arrays. Proc. Inst. Mech. Eng.Part C: J. -

Mech. Eng. Sci. 215: 1287–1301

[2] D’Alessio S J D and Dennis S C R 1994 A vorticity model

for viscous flow past a cylinder. Comput. Fluids 23:

279–293

[3] Dennis S C R and Chang G Z 1969 Numerical integration of

the Navier–Stokes equations for steady two-dimensional

flows. Phys. Fluids. Suppl. 12: II 88–II 93

[4] Dennis S C R and Young P J S 2003 Steady flow past an

elliptic cylinder inclined to the stream. J. Eng. Math. 47:

101–120

[5] Faruquee Z, Ting S K, Fartaj A, Barron R M and Carriveau R

2007 The effects of axis ratio on laminar fluid flow around an

elliptic cylinder. Int. J. Heat Fluid Flow 28: 1178–1189

[6] Fornberg B 1980 A numerical study of steady viscous flow

past a circular cylinder J. Fluid Mech. 98: 819–855

[7] Fornberg B 1985 Steady viscous flow past a circular cylinder

up to Reynolds number 600 J. Comput. Phys. 61: 297–320

[8] Grove A S, Shair F H, Peterson E E and Acrivos A 1964 An

experimental investigation of the steady separated flow past

a circular cylinder J. Fluid Mech. 19: 60–80

[9] Hunt J C R, Abell C J, Peterka J A and Woo H 1978

Kinematical studies of the flows around free or surface-

mounted obstacles; applying topology to flow visualization.

J. Fluid Mech. 86: 179–200

[10] Jackson C P 1987 A finite-element study on the onset of

vortex shedding in flow past variously shaped bodies.

J. Fluid Mech. 182: 23–45

[11] Johnson S A, Thompson M C and Hourigan K 2001 Flow

past elliptical cylinders at low Reynolds numbers. In:

Proceedings of the 14th Australasian Fluid Mechanics

Conference; Adelaide University, pp. 343–346

[12] Kumar D, Sourav K, Sen S and Yadav P K 2018 Steady

separation of flow from an inclined square cylinder with

sharp and rounded base. Comput. Fluids 171: 29–40

[13] Lighthill M J 1963 Boundary layer theory. In: Rosenhead L

(Ed.) Laminar Boundary Layers. Oxford University Press,

pp. 46–103

[14] Lugt H J and Haussling H J 1974 Laminar flow past an

abruptly accelerated elliptic cylinder at 45� incidence.

J. Fluid Mech. 65: 711–734

[15] Masliyah J H and Epstein N 1971 Steady symmetric flow past

elliptic cylinders. Ind. Eng. Chem. Fundam. 10: 293–299

[16] Saad Y and Schultz M H 1986 GMRES: A generalised

minimal residual algorithm for solving nonsymmetric linear

systems. SIAM J. Sci. Stat. Comput. 7: 856–869

[17] Sahin M and Owens R G 2004 A numerical investigation of

wall effects up to high blockage ratios on two-dimensional

flow past a confined circular cylinder. Phys. Fluids 16:

1305–1320

[18] Sen S and Mittal S 2017. A study on the far wake of elliptic

cylinders. Comput. Model. Eng. Sci. 113: 33–53

[19] Sen S, Mittal S and Biswas G 2009 Steady separated flow

past a circular cylinder at low Reynolds numbers. J. Fluid -

Mech. 620: 89–119

[20] Sen S, Mittal S and Biswas G 2012 Steady separated flow

past elliptic cylinders using a stabilized finite-element

method. Comput. Model. Eng. Sci. 86: 1–28

[21] Sivakumar P, Bharti R P and Chhabra R P 2007 Steady flow

of power-law fluids across an unconfined elliptic cylinder.

Chem. Eng. Sci. 62: 1682–1702

[22] Smith F T 1983 Interacting flow theory and trailing edge

separation � no stall. J. Fluid Mech. 131: 219–249

[23] Sourav K and Sen S 2017 On the response of a freely

vibrating thick elliptic cylinder of low mass ratio. J. Appl. -

Fluid Mech. 10: 899–913

[24] Sourav K, Kumar D and Sen S 2020 Vortex-induced vibrations

of an elliptic cylinder of low mass ratio: identification of new

response branches. Phys. Fluids 32: 023605-1–20

[25] Sugihara-Seki M 1993 The motion of an elliptic cylinder in

channel flow at low Reynolds numbers. J. Fluid Mech. 257:

575–596

[26] Tezduyar T E, Mittal S, Ray S E and Shih R 1992

Incompressible flow computations with stabilized bilinear

and linear equal-order-interpolation velocity pressure ele-

ments. Comput. Meth. Appl. Mech. Eng. 95: 221–242

[27] Yano H and Kieda A 1980 An approximate method for

solving two-dimensional low-Reynolds-number flow past

arbitrary cylindrical bodies. J. Fluid Mech. 97: 157–179

[28] Yoon D H, Yang K S and Choi C B 2010 Flow past a square

cylinder with an angle of incidence. Phys. Fluids 22,

043603-1–12

172 Page 18 of 18 Sådhanå (2020) 45:172