Sei sulla pagina 1di 9

Philippe Versailles1

e-mail: philippe.versailles@mail.mcgill.ca

Jeffrey M. Bergthorson
Associate Professor
Mem. ASME
e-mail: jeff.bergthorson@mcgill.ca
Department of Mechanical Engineering,
McGill University,
Montreal, QC, H3A 0C3, Canada

Optimized Laminar
Axisymmetrical Nozzle Design
Using a Numerically Validated
Thwaites Method
This paper presents the Thwaites method as an accurate and efficient design tool for laminar, axisymmetrical nozzles. Based on historical developments, it is improved to describe
internal flows with highly favorable pressure gradients in cylindrical coordinates. The
calculation of the core flow velocity distribution based on the continuity equation is proposed as a replacement to other sophisticated numerical methods. A remarkably good
agreement is obtained when comparing the results of the current Thwaites method
against those of computational fluid dynamics (CFD) simulations, for which the integral
boundary layer thicknesses are calculated with equations developed from first principles
in the course of the work. This consistency among the results and the low time and
resource costs of the Thwaites method confirm its applicability and usefulness as an engineering design and optimization tool. [DOI: 10.1115/1.4007155]

Introduction

Appropriately designed nozzles are critical for proper performance of large-scale wind tunnels [1], small-scale fundamental
combustion [2,3], or fluids experiments [4], as well as microscale
propulsion devices [5]. Over the years, various methods for
designing nozzle contractions have been used. They ranged from
by eye empirical techniques to others involving the selection of
a given velocity distribution along the axis, from which the stream
function, achieving a reasonable pressure gradient, is chosen as
the optimal nozzle contour. Moreover, panel and finite difference
methods provided more detailed inviscid (potential flow) solutions. These techniques are summarized in Ref. [6]. Bell and
Mehta [6,7] employed a 3D, inviscid, panel code to get the velocity distribution at the wall of contractions with noncircular cross
section that was then fed to the Thwaites method [8] for the prediction of the boundary layer characteristics. This removed the
need to rely on separation criteria, as was needed in earlier techniques. However, the work of Bell and Mehta did not analyze axisymmetrical internal flows, and the results of the numerical
method were solely validated against experimental data obtained
slightly downstream of the contraction outlet. Further studies
investigating the validity of the Thwaites method for the prediction of boundary layer characteristics inside nozzle contractions
are required. A similar methodology to the one of Bell and Mehta
has been used recently [9], with the potential flow solution
obtained using the OpenFoam CFD package, to compare the performance of transformed fifth-order polynomial shapes against the
original, well known, fifth-order profile proposed in Ref. [6]. Such
CFD and panel method computations are useful for case to case
comparison. Nevertheless, they are not practical for the study of
large design spaces, as often encountered in engineering situations, because they would require the generation of numerous 2D/
3D grids and a significant amount of time and resources in analyzing and postprocessing all the possible configurations.
In contrast to such detailed solutions, the Thwaites method has
been applied for the design of axisymmetrical nozzles, with the

potential core velocity distribution obtained simply from mass


conservation applied at each axial location [2,4]. The boundary
layer parameters were quickly calculated, which allowed a manual
adjustment of the remaining degrees of freedom of the 6th [4] and
7th [2] order polynomials in an attempt to minimize the possibility
of transition to turbulent flow. This methodology, even though it
yields well-performing nozzles, requires further validation.
Recently, Fujiso et al. [10] studied the flow in an axisymmetrical,
bicubical nozzle experimentally and numerically using the
Thwaites method with the velocity distribution calculated, presumably, as in Refs. [2] and [4]. However, the comparison of the
computations with experiments is again done at the outlet only
and, therefore, the accuracy of the Thwaites method inside the
contraction cannot be assessed. In addition, the experimental
boundary layer integral parameters at the outlet are evaluated with
equations for 2D planar flows, although Cipolla and Keith [11]
showed that they differ for axisymmetrical flows. Furthermore,
although the behavior of the integral parameters along the axial
direction obtained numerically in Ref. [10] is reasonable, the theoretical justification of this spatial variation is inexact. These previous studies collectively indicate the necessity of validating the
Thwaites method for axisymmetrical internal flows.
The current paper evaluates the accuracy of the Thwaites
method and demonstrates its usefulness as an engineering design
tool for axisymmetrical nozzles when used with a simple calculation of the potential core flow velocity as in Refs. [2] and [4].
First, the theory and development of the Thwaites method for cylindrical, internal flows with significant favorable pressure gradient is given. This is followed by descriptions of the algorithm
developed in the current work and of the benchmark CFD simulation along with its postprocessing into boundary layer integral parameters. The results from both methods are then compared in
detail in order to assess the accuracy of the Thwaites method. A
discussion on the utility of the Thwaites method in designing optimally performing nozzles follows.

2
1

Corresponding author.
Contributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received August 22, 2011; final
manuscript received July 10, 2012; published online September 27, 2012. Assoc.
Editor: Mark F. Tachie.

Journal of Fluids Engineering

The Thwaites Method

The Thwaites method is a semiempirical, one-dimensional tool


that predicts integral parameters of laminar boundary layers (i.e.,
d* and h) for given free-stream velocity distributions. It is based
on the Karman momentum integral equation (KMIE), Eq. (1),

C 2012 by ASME
Copyright V

OCTOBER 2012, Vol. 134 / 101203-1

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 11/25/2014 Terms of Use: http://asme.org/terms

which is developed from the NavierStokes equations simplified


for boundary layers (NSBL). They thus share the same assumptions: (1) the Reynolds number must be high and (2) the boundary
layer thickness sufficiently small. In Eq. (1), x is the axial coordinate, d* and h are the displacement and momentum thicknesses,
respectively, as given for external planar flows in Ref. [8] and for
external axisymmetrical flows in Ref. [11], U(x) is the free-stream
(potential) flow velocity, sw is the shear-stress at the wall, and q is
the density.
sw
dh
h dU d dU
2 


2
dx
U dx U dx
qU

(1)

The LHS term of Eq. (1) corresponds to the diffusive (viscous)


component of the NSBL, the third RHS term to the pressure component, and the remaining terms relate to the convective terms of
the NSBL. Multiplying Eq. (1) by (U  h)/,  being the kinematic
viscosity, inserting the HolsteinBohlen parameter k (Eq. (2a))
and the dimensionless wall friction parameter S (Eq. (2b)), and
assuming H  d*/h  H(k) gives Eq. (3).
h2  dU=dx


(2a)

sw  h
 Sk
qU

(2b)

U

 
d h2
 2  Sk  k  2 Hk Fk
dx 

(3)

Thwaites observed [12] that F(k) is well described by the linear


equation,
Fk  0:45  6:0  k

(4)

This function allows a straightforward resolution of the differential Eq. (3),


hx2 

0:45  

U6

U 5 dx

(5)

By performing the integral of Eq. (5) for a given U(x), the momentum thickness and the HolsteinBohlen parameter, k, are
obtained at every x-location. Thwaites proposed the fifth order
polynomial of k shown by the gray dash-dot line in Fig. 1 to determine H(k). This correlation is valid down to k 0.09, where
boundary layer separation is expected to occur. This separation
prediction capability of the Thwaites method is very useful in
assessing the performance of nozzles. Once the shape factor H is
obtained, the displacement thickness d* can be calculated and the
boundary layer is fully characterized as a function of x.
The range of validity of the Thwaites correlation for H(k) is restricted to k lower than 0.25, and inaccuracies in Eq. (4) have
been observed. Based on solutions to FalknerSkan velocity profiles, Dey and Narasimha [13] proposed the second order polynomial Eq. (6) in replacement to Eq. (4) and the H(k) correlation
shown by the gray dotted line in Fig. 1. Assuming that the second
order term of Eq. (6) is a small perturbation (DF) applied on the
original linear fit of Thwaites (Eq. (4)), they developed Eq. (7) for
the calculation of the momentum thickness. The first integral in
the bracket of the RHS is identical to the original solution of Eq.
(5), while the second is a usually small correction for highly
favorable pressure gradients. With these modifications, the
Thwaites method is expected to be valid for k up to 0.4, which
corresponds to highly accelerating flows [13].
Fk  0:45  6:0  k 2:0  k2
101203-2 / Vol. 134, OCTOBER 2012

(6)

Fig. 1 Correlations of the shape factor, H, against the


HolsteinBohlen parameter, k

0:45  
hx2 

U6

 x
0

U5 dx 0:9

x
0

dU
dx

2

1
 7
U

x0

!2
U 5 dx0

3
 dx5

(7)
For axisymmetrical bodies, Rott and Crabtree [14] developed
Eq. (9) based on Eq. (8), which is the cylindrical equivalent to
Eq. (3). The same H(k) and F(k) correlations as for external planar
flows are assumed to apply, boundary layer separation is again
anticipated at k 0.09, and an additional assumption is that the
boundary layer thickness d must be much smaller than the local
radius of the body r0(x). The expected accuracy of Eq. (9) is
65%.


U d r02  h2

 2  Sk  k  2 Hk Fk

r02 dx

0:45   x 2
hx2  2

r  U 5 dx
r0  U 6 0 0

(8)
(9)

In order to extend the range of validity of Eq. (9) to higher favorable pressure gradients, the strategy of Dey and Narasimha is
applied to axisymmetrical flows in the present work. Keeping
their F(k) correlation Eq. (6), a corrected equation for the momentum thickness in cylindrical coordinates is developed,
 x
0:45  

r02  U 5 dx
hx2  2
r0  U 6
0
3
!2
x0
x  2
dU
1
2
5
0
 2
r  U dx  dx5 (10)
0:9
r0  U7 0 0
0 dx
The H(k) correlations of Thwaites [12] and Dey and Narasimha
[13] are in close agreement for 0.1 < k < 0.17, but they display
significant differences for k < 0.1, as seen in Fig. 1. A new formula for H(k) is proposed here. To build the current H(k) relation
shown by the black solid line in Fig. 1, the correlation of Thwaites
is used for the low end of k values, owing to its historical widespread utilization, while the correlation of Dey and Narasimha is
used for the highest values of HolsteinBohlen parameter. The
union is at k  0.15, and efforts have been made to ensure that
H(k) is second-order continuous to obtain smooth solutions in
terms of displacement thickness.

3 CFD Computations and Comparison


With Thwaites Method
The same nozzle, intended to operate at a pressure of
2026.5 kPa (20 atm) and a temperature of 298 K, is analyzed using
Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 11/25/2014 Terms of Use: http://asme.org/terms

Fig. 2 Nozzle contour, first and second-order derivatives of


the radius as a function of x

the Thwaites method and CFD. Its contour is given by a 7th order
polynomial with first- and second-order derivatives equal to zero
at the inlet and outlet. The exit diameter of 0.01 m (0.400 ) is
selected based on previous well-performing nozzles [15], as well
as to minimize the required gas flow during the experiments. With
the inlet diameter of 0.032 m (1.2500 ), this leads to a suitable contraction ratio of 9.77 for flow relaminarization of disturbances [6].
The last two degrees of freedom of the polynomial are fixed by
adjusting the location of two points (x1, r0,1) and (x2, r0,2). The
nozzle has a length of 0.0762 m (300 ), which is sufficiently short to
minimize the growth of the boundary layer, but long enough to
prevent separation and instabilities in the concave section and to
achieve uniform flow at the outlet [2]. An entrance length corresponding to 20% of the radius at inlet of the contraction, r0(x 0),
is added upstream of the nozzle. This is the optimal distance
between the turbulence management assembly and the entrance of
the nozzle to prevent flow distortion through the last screen and to
minimize boundary layer growth [16]. Figure 2 shows the nozzle
contour without entrance length and the first- and second-order
derivatives of the radius r0 as a function of the axial location x.
3.1 Thwaites MethodIterative Calculations of the Boundary Layer Integral Parameters. In the current study, the
Thwaites method for axisymmetrical internal flows presented in
Sec. 2 is implemented in a Matlab algorithm. The core equation of
the code is Eq. (10), which, once solved, allows finding h(x),
d*(x), as well as other important flow properties. The nozzle
length is discretized with equally spaced nodes located at the
same axial position as for the CFD mesh for easier point-to-point
comparison. A supplementary length upstream of the simulated
nozzle has been added in the Thwaites simulation to account for
the momentum thickness at the nozzle inlet due to the growth of
the boundary layer in the settling chamber. The length is adjusted
such that the momentum thickness at x 0 (see Fig. 2) is the same
for both CFD and Thwaites simulations. In the end, the results of
the Thwaites code are obtained with 1384 equally spaced nodes
along the nozzle. However, preliminary results for a mesh consisting of 232 nodes almost perfectly overlap those obtained with the
finest discretization.
At first, U(x) must be determined in order to solve Eq. (10). In
contrast to previous studies using sophisticated panel or CFD
methods to obtain the potential core flow velocity [6,7,9], the current study assumes that the axial velocity is constant over the
cross section for all axial positions. The local velocity corresponds
to the volumetric flow rate normalized by the local area, and it can
be calculated with Eq. (11), derived from the continuity equation
for incompressible flows. The volumetric flow rate at one axial
position must be set in order to get U(x) at all other locations. For
that purpose, the outlet velocity U(xoutlet) is set to 3 m/s, a relevant
value for the current design.
Ux  r0 x2 Uxoutlet  r0 xoutlet 2

3.2 CFD Method. The CFD analysis of the nozzle shown in


Fig. 2 is performed with the ANSYS Fluent software. The implicit
formulation of the steady-state continuity and NavierStokes
equations is solved with the 2D, double-precision, axisymmetric,
segregated solver. The SIMPLEC pressure-velocity coupling
method is used and the equations are spatially discretized using a
second-order scheme. Based on preliminary Thwaites results, the
flow is assumed to be laminar throughout the nozzle. The working
fluid is air with constant density (23.68 kg/m3) and dynamic viscosity (1.8616  105 kg/m-s), both adjusted to match the design
operating condition of 2026.5 kPa (20 atm) and 298 K. The computational domain is shown in Fig. 3. No-slip and symmetry conditions are set on the wall and the axis, respectively. The nearly
flat axial velocity profile given by Eq. (12) is assumed to reproduce the flow coming out of a settling chamber [15]. In an attempt
to reproduce the real growth of the boundary layer, it is applied as
the inlet boundary condition instead of a constant velocity profile
that would lead to unrealistic nil values for d* and h and impact
the boundary layer behavior in the inlet section. A gauge static
pressure of 0.107 kPa (equal to the negative of the dynamic
pressure at the outlet) with respect to the operating pressure of
2026.5 kPa is specified at the outlet.
h

r i
uxinlet ; r uxinlet ; r 0  tanh 50  1 
R
h

i
r
0:306318  tanh 50  1 
m=s
0:015875 m
(12)
Six different meshes have been considered for the computation.
Their main properties are given in Table 1, where N and R

(11)

The solution to Eq. (10) is obtained by marching along the nozzle.


The three integrals are initially calculated between the first and
Journal of Fluids Engineering

second nodes using a first-order quadrature (trapezoidal rule) with


an error term O(Dx3). The momentum thickness at the second
node h(x2) is calculated along with k with Eq. (2a). The velocity
derivative in the HolsteinBohlen parameter is evaluated using a
first-order backward difference with an error term O(Dx). The
shape factor H is then obtained with the correlation shown by the
black solid line in Fig. 1 to find the displacement thickness at the
second node, d*(x2). Afterwards, the integrals of Eq. (10) are evaluated between the second and third nodes, again with the trapezoidal rule, and are added to the integrals between the first and
second nodes, resulting in integrals between the first and third
nodes. h, k, H, and d* are then found at x3. The procedure is
repeated for all nodes by marching toward the nozzle outlet.
d* corresponds to the effective displacement of the streamlines
in the potential flow due to the presence of the boundary layer.
The effect is analogous to a displacement of the wall by a distance
equal to d* into the flow. For the particular case of internal flows,
it results in a reduction of the effective radius (area) that will
increase the flow velocity U(x). Once d* is calculated for all axial
locations, it is then necessary to recalculate U(x) with Eq. (11),
but with the radius r0(x) replaced by the effective radius
reff (x) r0(x)  d*(x). With this new U(x), Eq. (10) is solved
again to obtain h, k, H, and d*. Another U(x) distribution is calculated, and the iterative process goes on until convergence of the
velocity: DU(x)/U(x) < 0.05% 8 x.

Fig. 3

Computational domain

OCTOBER 2012, Vol. 134 / 101203-3

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 11/25/2014 Terms of Use: http://asme.org/terms

Table 1 Characteristics of the CFD meshes

Mesh #

NI

RI

NA

RA

NO

RO

Total number
of cells

Mean difference in
h versus mesh #6

1
2
3
4
5
6

80
160
240
160
240
400

1.100
1.075
1.020
1.054
1.020
1.020

425
850
1275
1913
1950
1275

1
1
1
1
1
1

80
160
240
160
240
400

1.032
1.032
1.010
1.032
1.010
1.010

34,000
136,000
306,000
306,080
468,000
510,000

0.77%
0.36%
0.04%
0.29%
0.04%

represent the number of cells and the successive ratio of cell size
(the finest being at the wall), respectively. The subscripts I, A, and
O stand for the inlet, axis, and outlet edges, respectively, as shown
in Fig. 3. In order to assess mesh convergence, the mean difference over the length of the nozzle in terms of momentum thickness (h) between each grid and mesh #6 is included in the table.
Increasing the number of cells from 34,000 to 136,000 removed
oscillations in the boundary layer thicknesses sampled along the
axial direction and decreased the average difference by a factor
of 2. A smoother curve of H against x was obtained with mesh
#3, and the mean difference decreased by about an order of magnitude in comparison to grid #2. Increasing the number of nodes
on the axis, while keeping the total number of cells almost constant, yielded a higher mean difference for mesh #4. This is due to
the coarser mesh along the radial direction. On the other hand,
keeping the same mesh characteristics on the inlet and outlet
boundaries, but increasing the number of nodes on the axis with
mesh #5 did not change the mean difference in comparison to grid
#3. This indicates that 1275 nodes equally spaced on the axis are
sufficient. Furthermore, the relatively small difference in terms of
momentum thickness between meshes #3 and #6 does not require
the use of the finest mesh. Mesh #3 is then used to produce the
results presented in the remainder of this paper. In order to minimize the error, the CFD simulations were allowed to run until stabilization at minimal values of the scaled residuals of the
continuity, x-momentum, and y-momentum equations. Over most
of the 13,000 iterations required for convergence, the residuals
reduced smoothly and almost linearly (on a log scale) toward
1013, 1016, and 1017, respectively.
For the sake of brevity, images of the CFD solution will be
omitted. However, contours of the main variables (pressures,
velocities, vorticity, derivatives, etc.) have been verified to be free
of oscillations caused by insufficient mesh refinement. The flow
field consists of an accelerating potential core stream, within
which the vorticity is negligible and the total pressure constant,
surrounded by a very thin boundary layer close to the wall. In this
last region, the velocity decreases to zero at the wall due to the
no-slip condition, vorticity is much higher than in the core flow,
and the total pressure decreases toward the wall due to viscous
losses.

Fig. 4 Control volume for the calculation of the integral parameters in internal cylindrical coordinates

course of the present work. The 2D axisymmetrical control


volume used for this purpose is shown by the light gray area in
Fig. 4. It is bounded by the wall, by a surface coinciding with the
boundary layer height d, and by the two velocity profiles u(x,r)
and U0(x,r), the last being the profile that would be observed at
the same location as u(x,r) in absence of viscous effects. Hence,
the control volume is virtual, as it does not surround a physical
domain, but rather a theoretical one delimited by velocity profiles
observed in different circumstances.
As stated previously, the displacement thickness is the distance
by which the boundary layer deflects the streamlines of the potential core flow. It corresponds to the height of a layer of stagnant
fluid at the wall in the potential flow field U0(x,r) that would cause
the same reduction in mass flow rate as the boundary layer induces in the viscous flow (in comparison to the nonviscous flow).
The reduction of flow rate due to the boundary layer, equal to
m_ out , is given by the RHS term of Eq. (13). The equivalent reduction of flow rate in the stagnant layer of potential flow bounded by
the wall and d* is given by the LHS term of Eq. (13).
R
Rd

3.3 Boundary
Layer
Integral
Parameters
for
Axisymmetrical Internal Flows. Equations for d* and h for planar [8] and external axisymmetrical [11] flows have been developed based on the assumptions that (1) the deflection of the
streamlines due to the presence of the boundary layer does not
change the pressure field significantly in the potential flow and (2)
the velocity of the surrounding free-stream flow U is solely a
function of x. The first assumption has been verified for internal
nozzles with the CFD simulation described in Sec. 3.2. However,
the velocity in the nozzle core potential flow U0(x,r), analogous to
the free-stream velocity over a bluff-body, varies with the radial
location r, as illustrated by the gray velocity profile in Fig. 4.
Since U0 is a function of the radial location, a technique must be
derived to determine d* and h.
Exact equations for calculating the necessary boundary layer
thicknesses from CFD solutions have been developed in the
101203-4 / Vol. 134, OCTOBER 2012

q  U0 x; r  2pr  dr

q  U0 x; r  ux; r   2pr  dr

Rd

(13)
Since, U0(x,r) is equal to u(x,r) outside of the boundary layer,
Eq. (13) can be simplified to
Rd

U0 x; r  r  dr

ux; r  r  dr

(14)

which indicates that d* is specified to conserve the total flow rate


(RHS of Eq. (14)) in the inviscid flow. Similarly, the momentum
thickness h is the height of a layer of stagnant fluid at the wall that
would induce the same reduction of momentum flux in the potential flow (LHS of Eq. (15)) as the boundary layer itself causes in
the viscous flow (RHS of Eq. (15)). Performing a momentum balance analysis over the control volume of Fig. 4 to obtain the RHS
yields
Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 11/25/2014 Terms of Use: http://asme.org/terms

R
Rh

U02 x; r  r  dr

R
Rd

 2
U0 x; r  u2 x; r

 U0 x; R  d  U0 x; r  ux; r   r  dr
(15)
The first two terms in the bracket of the RHS are, respectively, the
inflowing and outflowing momentum fluxes through the left and
right boundaries of the control volume, respectively, while the
third term is the momentum flux exiting through the flow rate
m_ out . For ease of integration in the algorithm, Eq. (15) is rewritten
in a form similar to Eq. (14) by again invoking that U0 and u are
equal outside of the boundary layer,
Rh
0

U02 x; r  r  dr

 2
u x; r U0 x; R  d

 U0 x; r  ux; r   r  dr

(16)

The boundary layer thicknesses, d* and h, are taken as the location


where the LHS term of Eqs. (14) and (16) is equal to their
respective RHS term, which for a grid is a discrete position. Hence,
using Eqs. (14) and (16) leads to a step-like behavior of the integral
parameters against the axial location x. The rather fine mesh (#3)
used and the integration of a linear interpolation of d* and h
between two consecutive nodes minimize this step-like behavior in
the results. It is also necessary to select an appropriate velocity
profile U0, since the potential flow that would replace the boundary
layer region in the absence of shear stress is not readily obtainable
from the viscous CFD solutions. An approximate profile is devised
in the current study. As for the development of Eqs. (14) and (16),
the displacement of the streamlines by the boundary layer is
assumed to have a negligible effect on the pressure distribution in
the potential core flow. Therefore, U0(x,r) must be equal to u(x,r) in
this region. The height of the boundary layer d is taken as the location where the vorticity is three orders of magnitude smaller than
its maximum value at the wall. Within the boundary layer region,
U0(x,r) is calculated as

@u
U0 x; r > R  d ux; R  d 
 r  R  d (17)
@r rRd
Equation (17) states that U0 in the boundary layer region is linearly extrapolated from the axial velocity and its radial gradient
taken at the limit of the boundary layer (r R d).

Results

Figures 5 and 6 show the boundary layer thicknesses against


the axial location x as obtained from the Thwaites method and

Fig. 5 Momentum thickness versus axial location obtained


from the Thwaites method and CFD simulation

Journal of Fluids Engineering

Fig. 6 Displacement thickness versus axial location obtained


from the Thwaites method and CFD simulation

CFD. In these figures, the black solid and gray dashed lines are
the Thwaites and CFD solutions, respectively. The CFD solution
65% is also represented by dotted lines.
As seen in these figures, the agreement between both methods
is very good, within 5% everywhere in the nozzle except in the
last 7 millimeters, where the difference reaches 8.2% and 7.5%
for the momentum and displacement thicknesses, respectively.
This tends to confirm the expected accuracy (65%) of the
Thwaites method in cylindrical coordinates [8], specifically for internal flows. In addition, it validates Eqs. (10), (14), and (16)
developed during the work and demonstrates that the proposed
construction of U0(x,r), Eq. (17), is correct. The discrepancies in
the results might be due to the acceptable, but not exact, relation
for H(k) and F(k) in the Thwaites method, as well as to the u velocity profile that slightly varies radially in the CFD results, as
presented in the following. CFD simulations have been made with
an extension of 0.0254 m downstream of the nozzle in an attempt
to improve the agreement between both methods close to the outlet. However, this did not allow reconciliation of both Thwaites
method and CFD results.
The behavior of the boundary layer thicknesses along the nozzle can be explained from the evolution of the velocity derivative
(dU/dx) and of the favorable pressure gradient (@P/@x), shown
in Figs. 7 and 8, respectively. Exact solutions to the boundary
layer equations for accelerating free-stream velocities in 2D planar and axisymmetrical coordinates show that decreasing the pressure gradient @P/@x to more negative values (thus increasing the
flow acceleration and making the pressure gradient more favorable) reduces the boundary layer integral thicknesses [8,17].
In the first few millimeters of the nozzle, the pressure gradient
is relatively weak. This is because the flow acceleration caused by
the reduction of the radius, given by Eq. (18) for the Thwaites

Fig. 7 Free-stream velocity gradients versus axial location


obtained from the Thwaites method

OCTOBER 2012, Vol. 134 / 101203-5

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 11/25/2014 Terms of Use: http://asme.org/terms

Fig. 8 Negative of the pressure gradient versus axial location


obtained from the Thwaites method and CFD simulation

Fig. 10 Shape factor versus k obtained from the Thwaites


method and CFD simulation

method and showed by the gray dashed curve in Fig. 7, is relatively small, since r0 (x 0) 0. The growth of the boundary layer
results in a small acceleration of the flow and is the dominant
mechanism of flow acceleration in the entrance length and the first
few millimeters of the nozzle, as shown by the black dash curve
in Fig. 7.

dU
2 dr0
Uxoutlet  r0 xoutlet 2  3 
(18)
dx due to
r0 dx

dial variation of the u velocity profile (CFD) in contrast to the


hypothesized flat profile in the Thwaites method) likely impact
the agreement. This potential source of error remains to be
studied.
The close agreement between the integral parameters shown in
Figs. 5 and 6 also indicates that the shape factor H is properly estimated in the Thwaites method. It has been verified that both methods predict values less than 7.5% apart from each other over the
whole domain. This is because the H(k) correlation used in
the Thwaites code tends to reproduce the one calculated from the
CFD (with k calculated by means of a first-order forward derivative
of the axial velocity evaluated at the edge of the boundary layer), as
shown in Fig. 10. The bumps on the lower part of the CFD curve
for k > 0.125 are due to the limited order of the interpolation function in the Matlab postprocessing algorithm. In the course of the
analysis, it has been noticed that the main discrepancies in terms of
H(k) attaining 6.5% are observed in the convex section of the
nozzle. The reason for the inverse shape curve obtained with CFD
has yet to be investigated, but the results suggest that it is somehow
related to the curvature of the nozzle contour. Even if the H(k) correlation used in the Thwaites code is reasonable and yields nice
results (see Figs. 5 and 6), it could still be improved for flows going
through concave/convex nozzles. This would likely yield to a better
agreement between the Thwaites method and CFD, because H(k) is
involved in the determination of d*, which impacts U(x), which in
turn influences k and h. The loop of interactions between the parameters of the Thwaites method is illustrated in Fig. 11.
To further confirm the validity of the current Thwaites code,
the F(k) relation, Eq. (6), is plotted in Fig. 12 against values
obtained from CFD, with F and k evaluated at the edge of the
boundary layer (r R  d). The agreement, even though not perfect, is relatively good. The Thwaites correlation captures the negative slope of both branches of the CFD relation for F(k). The
cause of this dual-branched curve remains to be determined, but it
is expected that a more complex and accurate F(k) in the development of Eq. (10) would allow more accurate reproduction of the
CFD results (see Fig. 11). However, this would be at the expense
of a more problematical transformation of Eq. (8) into Eq. (10).

radius

As x rises, the radius change of the nozzle is dominant in determining dU/dx (solid line in Fig. 7), and the increasing favorable
pressure gradient results in a decrease of the boundary layer thicknesses along the x-direction, as seen in Figs. 5 and 6. This reduction of the integral parameters is also enhanced in the concave
region, which extends up to x 0.0523 m. It has been shown that,
for a given free-stream velocity with imposed velocity gradient,
increasing the curvature of a 2D plate reduces the boundary layer
thicknesses significantly [18]. Around x 0.066 m, the favorable
pressure gradient reaches its maximum value (highest @P/@x)
and then decreases toward the outlet of the nozzle because of a
significant reduction in the absolute value of r00 (x) in this region.
As the pressure gradient decreases, its coercive effect on the
growth of the boundary layer reduces and d* and h increase toward the nozzle outlet.
The agreement of the Thwaites method with CFD is partly due
to the consistency between U and u, and to the fact that the viscous velocity profile has only a small variation over the radius in
the core potential region at a given axial location. This can be
seen in Fig. 9, which shows that u(x, r 0) is close to u(x,
r R  d) almost everywhere in the nozzle. However, even the
small difference between u(x, r 0) and u(x, r R  d) (i.e., ra-

4.1 Flow Laminarity. The Thwaites method is valid for laminar flows, and the motive of the present paper is to demonstrate

Fig. 9 Axial velocity versus axial location obtained from the


Thwaites method and CFD simulation at different radial
positions

101203-6 / Vol. 134, OCTOBER 2012

Fig. 11 Interaction between the parameters of the Thwaites


method

Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 11/25/2014 Terms of Use: http://asme.org/terms

Fig. 12 F(k) correlation versus k obtained from the Thwaites


method and CFD simulation

its usefulness in the design of laminar nozzles. For these reasons,


criteria accounting for (1) the natural transition of the boundary
layer and (2) the transition due to the formation of Gortler vortices
in the concave section of the nozzle are applied to the current
design.
The first type of transition is characterized by the Orr
Sommerfeld equation. Its resolution for FalknerSkan flows (2D
flat wedge) with and without heating and for various pressure gradients, as well as for the stream over a flat plate with suction and
injection through the wall, showed that the critical d*-based Reynolds number for transition is a universal function of the shape
factor H, as shown in Fig. 13 and reported by Wazzan et al. [19].
Although not widely validated, this relationship between Red*crit
and H is proposed as a useful engineering tool to determine the
stability of the boundary-layer without requiring the laborious resolution of the OrrSommerfeld equation.
In the current study, the data points [19] shown in Fig. 13 are
described by a best-fit polynomial. For both computation methods,
this fit was used to obtain the critical Reynolds number. Natural
transition is expected to be prevented if the actual Reynolds number, Red*, is smaller than the critical one everywhere in the nozzle.
This is verified for the current nozzle with a maximum value of
Red*/Red*crit lower than 0.5, as shown in Fig. 14. It is observed by
comparing Fig. 14 to Figs. 5 and 6 that the critical regions in
terms of high Reynolds number ratio correspond to zones of growing boundary layer. The agreement is again quite good, with the
discrepancies in Fig. 14 explained by an underestimation of d*
close to the inlet and an overestimation of H close to the outlet by
the Thwaites algorithm in comparison to the CFD simulation.
The transition of the boundary layer can also be forced by
streamwise-oriented, counter-rotating vortices, also known as
Gortler vortices, that can form in the concave section of nozzles.

Fig. 14 Red*/Red*crit ratio versus axial location obtained from


the Thwaites method and CFD simulation

Fig. 15 Gortler parameter versus axial location obtained from


the Thwaites method and CFD simulation

A review on this topic is given in Ref. [20]. Based on theoretical


stability analysis, it has been reported that these vortices form
only over a threshold value of the Gortler parameter, G, defined
by Eq. (19) [1]. To prevent the formation of these vortices, G
must be minimized and kept below 53 [2,4], as experimentally
measured in Ref. [1].
G x Re2h  h 

d 2 r0
dx2

(19)

The performance of the current nozzle in terms of forced transition is shown in Fig. 15 for both computational methods. Again, a
good overlap of the curves is obtained, and the Gortler parameter
is well below the critical value of 53. Therefore, Gortler vortices
should not form for the operating condition studied, and forced
transition of the boundary layer is not expected.

5 The Thwaites Method as an Engineering Design


Tool for Nozzles

Fig. 13 Displacement-thickness-based critical Reynolds


number for transition as a function of the shape factor H

Journal of Fluids Engineering

A major advantage of the Thwaites method, apart from its short


computing time and good accuracy in comparison to CFD, is its
ease of integration in an optimization loop. Many contours,
described by polynomials of various orders, having different
lengths, inlet, and outlet diameters, can be analyzed within a few
hours with the current Thwaites algorithm. This makes it a useful,
simple, and efficient engineering design tool for 2D axisymmetrical nozzles.
In the present study, the diameters and length of the nozzle are
selected based on design requirements prior to optimizing its
shape. Therefore, only the points (x1,r0,1) and (x2,r0,2) (see Sec. 3),
fixing the last degrees of freedom of the polynomial, are varied.
Among all the polynomials calculated, only the ones for which
dr0/dx is lower than or equal to 0 everywhere in the domain are
OCTOBER 2012, Vol. 134 / 101203-7

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 11/25/2014 Terms of Use: http://asme.org/terms

analyzed, since nozzles with positive radius derivative would be


likely to suffer boundary layer separation. A few million randomly generated nozzle contours respect this criterion and are analyzed with the Thwaites algorithm presented and validated in this
paper. Since the objective of the iteration loop is to obtain the nozzle least prone to boundary layer transition, the various contours
are compared based on their maximum values of Red*/Red*crit and
Gortler parameters. The nozzle achieving the best performance in
terms of natural transition (lowest maximum value of Red*/
Red*crit) has a Gortler parameter G about an order of magnitude
higher than the one having the best characteristics with regard to
the Gortler vortices. However, this last profile has a maximum
value of Red*/Red*crit 8 x only very slightly higher than the former.
For these reasons, the profile achieving the lowest G parameter is
selected as the optimal profile for the current length, inlet, and
outlet diameters of the nozzle. This contour, presented in Fig. 2,
has been thoroughly analyzed in the present paper. This nozzle
design is expected to achieve the necessary laminar flow conditions at high pressures.

Conclusion

This paper demonstrates the applicability of the Thwaites


method as a design tool for laminar, axisymmetrical nozzles. A
novel equation predicting the momentum thickness in cylindrical
coordinates for conditions ranging from unfavorable to highly
favorable pressure gradients has been proposed along with a new
correlation for the shape factor H. The core flow velocity distribution required by the method was determined from a simple mass
balance analysis. After convergence, the boundary layer integral
parameters, pressure gradient along the axis, and core flow velocity obtained by the method showed great agreement with CFD
simulations performed at the exact same operating condition. This
confirms the expected accuracy of the Thwaites method in cylindrical coordinates and verifies its applicability to axisymmetrical
internal flows. It also validates the proposed calculation technique
of the core flow velocity that removes the need for finer computation methods. In addition, agreement in terms of natural and
forced boundary layer transition criteria has been observed
between the two methods, and the laminar behavior of the flow,
required for the Thwaites method to be valid, is verified. It also
indicates the attainment of laminar flow conditions at 2026.5 kPa
(20 atm), which was the aim of the current design effort. The accuracy of the Thwaites method developed here and its low computational cost make it a useful engineering design tool for
axisymmetrical nozzles.
To obtain the integral parameters from CFD simulations, new
equations for the boundary layer displacement and momentum
thicknesses for axisymmetrical internal flows have been derived
from first principles. A construction method for the inviscid velocity profile involved in these equations has also been presented,
and the resulting agreement between both methods demonstrates
their validity.
The current work focused on HolsteinBohlen parameter values
ranging from approximately 0 to 0.2, thus for relatively mild
favorable pressure gradients. A validation of the Thwaites
method for more aggressive favorable pressure gradients, with
HolsteinBohlen parameter values up to 0.4, is left for future studies. Experimental confirmation of the results presented here would
also be a valuable contribution, along with an analysis of the potential sources of error in the Thwaites method. Comparison of the
method with CFD at other operating pressures, even though not
anticipated to negatively affect the agreement, should also be
conducted.

Acknowledgment
The authors are thankful to Paul E. Dimotakis, John K. Northrop Professor of Aeronautics and Professor of Applied Physics at
the California Institute of Technology, for proposing this nozzle
101203-8 / Vol. 134, OCTOBER 2012

design methodology and for helpful discussions during the early


work on this problem. The support of P. Versailles by the Natural
Sciences and Engineering Research Council of Canada (NSERC)
through an Alexander Graham Bell Canada Graduate Scholarship
and McGill University by means of a Total-ISEAD/Brace Water
Resources Fund/McGill Engineering Doctoral Award is gratefully
acknowledged.

Nomenclature
F
G
H
KMIE
LHS
m_
N
NSBL
P
R
r
r0
r00
r000
reff
Red*
Red*crit
Reh
RHS
S
u
U(x)
U0(x,r)
x

dimensionless parameter defined in Eqs. (3) and (6)


Gortler parameter
shape factor
Karman momentum integral equation
left-hand-side term
mass flow rate
number of cells
NavierStokes equations simplified for boundary layers
static pressure
successive ratio of cell size
radial location
local radius of the nozzle
first-derivative of the local radius of the nozzle
second-derivative of the local radius of the nozzle
effective radius (r0 d*)
Reynolds number based on the displacement thickness
critical Reynolds number based on the displacement
thickness
Reynolds number based on the momentum thickness
right-hand-side term
dimensionless wall friction parameter
local axial velocity
free-stream flow velocity over a body, or alternatively
constant, area averaged, velocity in the core flow
potential (unaffected by viscous force) flow in the
nozzle
axial location

Greek Symbols
d
d*
h
k

q
sw

boundary layer thickness


displacement thickness
momentum thickness
HolsteinBohlen parameter
kinematic viscosity
gas density
shear stress at the wall

Subscripts
A on the axis
I at the inlet
O at the outlet

References
[1] Liepmann, H. W., 1943, Investigations on Laminar Boundary-Layer Stability and Transition on Curved Boundaries, National Advisory Committee
for Aeronautics, Technical Report No. NACA-WR-W-107, NACA-ACR3H30.
[2] Bergthorson, J. M., 2005, Experiments and Modeling of Impinging Jets and
Premixed Hydrocarbon Stagnation Flames, Ph.D. thesis, California Institute of
Technology, Pasadena, CA.
[3] Bergthorson, J. M., Salusbury, S. D., and Dimotakis, P. E., 2011, Experiments
and Modelling of Premixed Laminar Stagnation Flame Hydrodynamics,
J. Fluid Mech., 681, pp. 340369.
[4] Dowling, D. R., 1988, Mixing in Gas Phase Turbulent Jets, Ph.D. thesis, California Institute of Technology, Pasadena, CA.
[5] Hao, P.-F., Ding, Y.-T., Yao, Z.-H., He, F., and Zhu, K.-Q., 2005, Size Effect
on Gas Flow in Micro Nozzles, J. Micromech. Microeng., 15(11), pp.
20692073.
[6] Bell, J. H., and Mehta, R. D., 1988, Contraction Design for Small Low-Speed
Wind Tunnels, NASAAmes Research Center/Stanford University, Technical Report No. NASA-CR-177488.
[7] Bell, J. H., and Mehta, R. D., 1989, Boundary-Layer Predictions for Small
Low-Speed Contractions, AIAA J., 27(3), pp. 372374.

Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 11/25/2014 Terms of Use: http://asme.org/terms

[8] White, F. M., 1991, Viscous Fluid Flow, McGraw-Hill, New York.
[9] Doolan, C. J., 2007, Numerical Evaluation of Contemporary Low-Speed
Wind Tunnel Contraction Designs, ASME J. Fluids Eng., 129(9), pp.
12411244.
[10] Fujiso, Y., Wu, B., Grandchamp, X., and Hirtum, A. V., 2011, Caracterisation
` La Validation Experimentale De
Dun Convergent Axisymetrique Destine A
Production De Parole, Actes des 9e`mes Rencontres Jeunes Chercheurs en Parole, pp. 14.
[11] Cipolla, K. M., and Keith, W. L., 2003, Momentum Thickness Measurements
for Thick Axisymmetric Turbulent Boundary Layers, ASME J. Fluids Eng.,
125(3), pp. 569575.
[12] Thwaites, B., 1949, Approximate Calculation of Laminar Boundary Layer,
Aeronaut. Q., 1(Part 3), pp. 245280.
[13] Dey, J., and Narasimha, R., 1990, An Extension of the Thwaites Method for
Calculation of Incompressible Laminar Boundary Layers, J. Indian Inst. Sci.,
70(1), pp. 111. Available at http://journal.library.iisc.ernet.in/archives/v9-79/
Vol.70(1-6)1990/Vol.70(1)1990/vol70(1)1990.pdf

Journal of Fluids Engineering

[14] Rott, N., and Crabtree, L. F., 1952, Simplified Laminar Boundary-Layer Calculations for Bodies of Revolution and for Yawed Wings, J. Aeronaut. Sci.,
19(8), pp. 553565.
[15] Bergthorson, J. M., Sone, K., Mattner, T. W., Dimotakis, P. E., Goodwin, D.
G., and Meiron, D. I., 2005, Impinging Laminar Jets at Moderate Reynolds
Numbers and Separation Distances, Phys. Rev. E, 72(6), p. 066307.
[16] Mehta, R. D., and Bradshaw, P., 1979, Design Rules for Small Low Speed
Wind Tunnels, J. R. Aeronaut. Soc., 83(827), pp. 443449.
[17] Schlichting, H., 1979, Boundary-Layer Theory, McGraw-Hill, New York.
[18] Ozalp, A. A., and Umur, H., 2003, An Experimental Investigation of the Combined Effects of Surface Curvature and Streamwise Pressure Gradients Both in
Laminar and Turbulent Flows, Heat Mass Transfer, 39(10), pp. 869876.
[19] Wazzan, A. R., Gazley, C., Jr., and Smith, A. M. O., 1979, Tollmien-Schlichting Waves and Transition: Heated and Adiabatic Wedge Flows With Application to Bodies of Revolution, Prog. Aerosp. Sci., 18, pp. 351392.
[20] Saric, W. S., 1994, Gortler Vortices, Annu. Rev. Fluid Mech., 26(1), pp.
379409.

OCTOBER 2012, Vol. 134 / 101203-9

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 11/25/2014 Terms of Use: http://asme.org/terms

Potrebbero piacerti anche