You are on page 1of 21

American Institute of Aeronautics and Astronautics

1
Approximate Solution of a Laminar Jet Discharged into a
Dead End Tube
John M. Crane,
*
Yedidia Neumeier,

Prasad Bhave,

and Ben T. Zinn

Georgia Institute of Technology, Atlanta, Georgia, 30332


An approximate analytical solution of a laminar jet discharged into an infinitely long
cylindrical cavity with a dead end is developed. The employed method assumed velocity
profiles whose parameters were determined by solving the integral forms of the continuity,
momentum, and energy equations. A novel feature of the method is the use of calculus of
variations to obtain a minimum dissipation solution. The solution predicted a few
fundamental characteristics of such a jet. First, the jet expands asymptotically to a finite
width, which is 52% of the cavity width and is independent of the initial jet radius. Second,
after the jet is fully expanded, the centerline velocity decays nearly linearly and reaches
stagnation. Thus, the model predicted that in a realistic cavity with a finite length far larger
than the theoretical jet penetration, the flow stagnates upstream of the dead end and does
not turn around at the closed end. Significantly, the model predicted that with proper
scaling, the behavior of the jet for various radii and jet velocities can be collapsed to a single
characteristic. The model predictions were compared to CFD simulations and experimental
measurements and were found to be in good agreement.
Nomenclature
A = System matrix
B = System vector
, , a b c = Coefficients
F
= Radial shape function
f
= Ratio of forward and backward velocity
J = Momentum function
L = Cavity length
p
L = Jet penetration length
n = Exponent
p = Pressure
Re = Reynolds number
R = Radius
u = X component of the velocity
v = Y component of the velocity
Greek
= Radius ratio
= Dissipation function
= Normalized radial distance
= Jet width
= Viscosity
= Density
= Kinematic viscosity
= Dissipation functional
Scripts
[ ]
$
= Normalized variable
[ ]
asy
= Asymptote
[ ]
B
= Reverse direction
[ ]
F
= Forward direction
[ ]
i
= Related to the inner tube
[ ]
ii
= Normalized by inner tube radius
[ ]
io
= Normalized by inner and outer tube radii
[ ]
max
= Maximum
[ ]
o
= Related to the outer tube
[ ]
oo
= Normalized by outer tube radius
*
Graduate Research Assistant, School of Aerospace Engineering, AIAA Member

Adjunct Professor, School of Aerospace Engineering, AIAA Member

Graduate Research Assistant, School of Aerospace Engineering, AIAA Member

David S. Lewis Jr. Chair and Regents Professor, School of Aerospace Engineering, AIAA Fellow
47th AIAA Aerospace Sciences Meeting Including The New Horizons Forum and Aerospace Exposition
5 - 8 January 2009, Orlando, Florida
AIAA 2009-387
Copyright 2009 by J. M. Crane, Y. Neumeier, and B. T. Zinn. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
American Institute of Aeronautics and Astronautics
2
I. Introduction
his paper describes an approximate solution for a laminar jet discharged into a long tube.
**
The motivation
behind this type of flow is the recently developed Stagnation Point Reverse Flow (SPRF) combustor
1
whose
basic features are shown in Fig. 1. In contrast to state of the art combustors where the reactants and products enter
and leave the combustor through opposite ends, the reactants and products enter and leave the SPRF combustor
through the same plane opposite a closed end. Specifically, the fuel and air are supplied to the SPRF combustor
through concentric tubes located at the center of the open end. The flame produces a stream of hot combustion
products that move toward the closed end of the combustor where they are decelerated (in the stagnation zone) to
near zero velocity and forced to turn around and flow out of the combustor through an annular reverse flow stream.
Thus, the combustor contains counter-flowing streams of outgoing products and incoming reactants that come into
contact in a shear layer where the hot products and radicals mix with the reactants (see the left schematic in Fig. 1).
The combustion process in this combustor depends heavily upon the interaction of these two counter-flowing
streams. Therefore, basic understanding of the flow of a jet into a dead end tube is an essential step toward
understanding the combustion process in the SPRF combustor.
The problem of a jet discharged into an axisymmetric dead-end channel was discussed by Abramovich,
2
who
considered an incompressible turbulent jet discharged into a finite tube. Abramovich assumed that the turn of the jet
occured at the dead end where the fluid was treated as ideal. Amano
3
performed a numerical (CFD) study of
turbulent axisymmetric jets flowing into closed tubes. Most of his results were obtained for relatively short tubes
having length to injector diameter ratios between two and five. However, a case with a much longer tube was also
presented with a length to injector diameter ratio of 52 and tube to injector diameter ratio of 5.3. For this case, it was
noted that the centerline velocity decayed to zero near the midpoint of the tube. He concluded that the effect of the
inlet flow propagated only to a certain length for a very long closed tube. This finding suggests that with a long
enough tube, the flow will eventually stagnate irrelevant of the existence of a dead end. Eckmann et. al.
4
measured
the flow of a laminar jet into a dead end tube in the context of the flow in an artificial lung.
5
They investigated jet
penetration over a range of Reynolds numbers from 50 to 400 and inlet to exit tube diameter ratios of 0.1 to 0.3
**
The meaning of a long tube will be discussed in the problem formulation.
T
Combustor
Wall
Fuel & air
injector
Reaction
zone
Reaction
zone
Shear Layer
Air/Products
Mixing
Hot
Products
Air
Fuel
Shear Layer
Air/Fuel
Mixing
-
Combustor
Wall
Flow Features of the
Upper Region of Combustor
Figure 1. A schematic of the SPRF combustor (center) with the flow features of the upper region of the
combustor (left). Images of the combustor burning natural gas in both premixed and non-premixed modes of
operation (right).
Premixed Non-Premixed
American Institute of Aeronautics and Astronautics
3
using flow visualization and particle image velocimetry. Due to the similarity of the flow characteristics and
geometry, their measurements were used to validate the predictions of the laminar jet model described herein.
II. Theory
A. Problem Formulation
To put the fluid mechanics of the discussed problem in context, consider the three limiting cases of a jet
discharged into a long tube shown in Fig. 2. The left case shows the flow features in a closed tube with all viscous
forces null. In this case, the jet penetrates the full distance of the tube, turning around near the closed end in a
potential flow. The center case shows the flow features in a long open tube taking viscous forces in the fluid into
account while neglecting shear forces at the wall. In this case, the flow would emerge at the far end with a fully
developed velocity profile. In the absence of wall shear forces, the cross sectional integral of the momentum at the
inlet and outlet is equal. The right case takes both fluid viscous forces and shear forces at the wall into account. As
the incoming flow proceeds into the tube, its momentum is dissipated by the wall forces. If the tube is just long
enough, the shear forces at the wall dissipate the momentum of the incoming jet, diminishing the mass flux through
the cross sectional area. For such a long tube, the flow must reverse itself and exit near the injector. Solving this
case is the focus of this study.
Figure 2. Three limiting cases of a jet discharged into a long tube: inviscid potential flow (left), viscous flow
with no wall shear (center), and viscous flow with wall shear (right).
Consider now the flow profile whose axial velocity component u is shown in Fig. 3. The boundary conditions
for the axial and radial velocities, u and v , respectively, are
0
0
r
u
r
=
c
=
c
(1)

0
0
o
r r R
v v
= =
= = (2)
0
r
u
=
= (3)
0
o
r R
u
=
= . (4)

American Institute of Aeronautics and Astronautics
4
It is important to note that there are no end boundary conditions since
the end is some arbitrarily long distance from the inlet, and in fact, as
previously discussed, may even be open.
To simplify the problem, low Mach number (thus constant density)
flow is assumed. Further, we assume that the pressure is radially (but
not axially) uniform. Applying continuity and axial momentum
conservation in cylindrical coordinates gives
1 ( )
0
u rv
x r r
c c
+ =
c c
(5)

1 1 u u u p
u v r
x r r r r x
c c c c c | |
+ =
|
c c c c c
\ .
v (6)

where v is the kinematic viscosity.
Since all flow enters and exits through the same plane, the net flow
rate through any cross section is null. The continuity integral equation
is thus
0
( ) 0
o
R
u r r dr =
}
. (7)

Integrating Eq. (6) in the radial direction and using the continuity Eq.
(5) yields
2
2
0
1
2
O
o
R
o
o
r R
R u dp
u rdr R
x r dx
=
c c
=
c c
}
v . (8)

It should be noted that the integral of the radial momentum equation
with uniform radial pressure distribution would provide no new information, thus is not used.
Next, the energy equation is derived. A detailed derivation of the equation is given in Appendix A with the main
steps outlined below. First, both sides of Eq. (6) are multiplied by the product ur and, using Eq. (5), it is
manipulated to
( ) ( )
2
3 2
1 1 1
2 2
u u p
ru rvu r ru ur
x r r r r x
c c c c c c | | | |
+ = +
| |
c c c c c c
\ . \ .
v v . (9)

Integrating both sides of Eq. (9) with respect to r gives
2
3
0 0
1
2
O o
R R
u
u rdr rdr
x r
c c | |
=
|
c c
\ .
} }
v . (10)
Note that since the axial derivative of the pressure is assumed radially uniform, the pressure term disappears after
integrating Eq. (9), since the integral of the other part of that term, ur , is null from continuity. Also absent is the
contribution of the wall shear stress force. Both of these terms are accounted for in the integral momentum Eq. (8).
Since the integral energy Eq. (10) is derived from the continuity and momentum equations, the use of all three
Eqs. (7), (8), and (10), is apparently superfluous. To correct this impression, consider the following. Assume that the
flow field is divided into two regions containing the forward and reverse flow. Solving such a problem would
require applying the integral form of the continuity and momentum equations to each domain, providing four
equations. Alternatively, the problem could be formulated with continuity and momentum in one domain and
continuity and energy integrated over the entire cross section resulting in the same number of equations (four). This
arrangement takes advantage of the null flow rate over a cross section and the eliminated pressure term in the energy
equation. It is thus evident that a proper solution may use a combination of the different equations without over
R
i
r

R
o
x
Figure 3. Schematic of the flow and the
geometry.
American Institute of Aeronautics and Astronautics
5
defining the problem. As a side note, while this problem is nicely divided into two domains, one could divide a
problem into any number of domains and apply the governing equations to each. This would provide for an even
larger number of possible combinations of the governing equations.
B. Solution approach
As mentioned above, the flow field is divided into two regions containing the incoming (forward) and outgoing
(backward) flow. The velocity distribution can be described as a combination of two profiles
( ) ( ) :
( ) ( ) :
F F
approx
B B o
o
r
u x F r
u
r
u x F R r
R
q q

q q

<

> >

(11)
where ( )
F
u x and ( )
B
u x are cross-sectional velocity magnitudes, q is normalized radial distance,
F
F and
B
F are
radial functions yet to be defined, and ( ) x is the jet spread, as shown in Figure 3. The boundary conditions given
in Eqs. (1)-(4) impose
0
0
F
dF
d
q
q
=
= (12)
1 1
0
F B
F F
q q = =
= = . (13)
The following non-dimensional quantities are introduced
$
o
R

(14)
0

F
x
u
u
u
=
(15)
0
Re
F o
x
o
u R
=

v
(16)
0
Re
F i
x
i
u R
=

v
(17)
i
o
R
R
c (18)
$
Re
oo
o o
x
x
R
(19)
$ $
$
Re Re
oo
oi io
o i i o
x x x
x x
R R c
= = = (20)
American Institute of Aeronautics and Astronautics
6
$
$
2
Re
oo
ii
i i
x x
x
R c
= (21)
0

F
x
p
p
u
=
(22)
0

=
. (23)
Note the three distance scales, Eqs. (19)-(21), that arise from the possible combination of the radii and
corresponding Reynolds numbers. Later, the merits of each length scale will be discussed. Substituting Eq. (11) into
Eqs. (7), (8), and (10) along with the normalized quantities, Eqs. (14)-(23), the following three equations are
obtained.
The continuity equation transforms to
$ $ $ $
( )
$ $ $
( )
2 2
1 2 3
1 1 0 F B B a u a u a u + + = , (24)
the energy equation becomes
$ $ $ $
( )
$ $ $
( )
$ $ $
$
$
2 3 2 3 3 2 2 2
1 2 3 4 5 6
1
1 1
2
1
F B B F B B
oo
d
b u b u b u b u b u b u
dx

| |
| |
+ + = + + | |
|
\ .
\ .
, (25)
and the momentum equation converts to
$ $ $ $
( )
$ $ $
( )
$
( )
$

2 2 2 2 2
1 2 3
1
1 1
1 1

2 1
momentumflux
B
F B B B
oo oo
dF d d p
c u c u c u u
dx d dx
J
q

q

=
| |
+ + =
|
\ .

1444444442444444443
(26)
where the coefficients are given by
1 1 1
1 2 3
0 0 0
, ,
F B B
a F d a F d a F d q q q q q = = =
} } }
(27)
1 1 1
3 3 3
1 2 3
0 0 0
, ,
F B B
b F d b F d b F d q q q q q = = =
} } }
(28)
2 2 2
1 1 1
4 5 6
0 0 0
, ,
F B B
dF dF dF
b d b d b d
d d d
q q q q q
q q q
| | | | | |
= = =
| | |
\ . \ . \ .
} } }
(29)
1 1 1
2 2 2
1 2 3
0 0 0
, ,
F B B
c F d c F d c F d q q q q q = = =
} } }
. (30)
These coefficients express the various moments of the spatial functions,
F
F and
B
F . Equations (24) through (26)
can be presented in compact form
$
$
$
[ ] [ ]

F
B
oo
u
d
A u B
dx




=
`


)
(31)
where the elements of the A matrix and B vector are given in the table below.
American Institute of Aeronautics and Astronautics
7
Matrix Elements Vector Elements
$
2
11 1
A a =
1
0 B =
$
( )
$ $
( )
2
12 2 3
1 1 A a a = +
$ $ $
$
$
2 2 2
2 4 5 6
1
F B B B b u b u b u

| |
= + + |
|

\ .
$ $ $ $
( )
$ $
( ) 13 1 2 3
2 2 1 1 2 F B B A a u a u a u = +
$
$
3
1
1
2
1
B
B
oo
dF u dp
B
d dx
q
q

=
=

$ $
2 2
21 1
3
2
F A b u =
$ $
( )
$
( )
$
( )
2
22 2 3
3
1 1
2
B A u b b = +
$ $ $ $
( )
$ $
( )
3 3 3
23 1 2 3
1
1 1 2
2
F B B A b u b u b u = +
$ $
2
31 1
2 F A c u =
$ $
( )
$ $ $
( )
2
32 2 3
2 1 2 1 B B A c u c u = +
$ $ $
( )
$ $
( )
2 2
33 1 2 3
2 2 1 1 2 F B B A c u c u c u = +
These three equations (Eqs. (24)-(26)) contain four unknowns: the two velocity amplitudes,
F
u and
B
u , the jet
spread,

, and the pressure, p . This imbalance between the number of equations and unknowns comes not from
the way the equations were chosen, but rather from the infinite cavity length. Consequently, the equations developed
are not subjected to a boundary condition at the closed end. Instead, an additional constraint is required for closure
of the problem.
To provide such closure, an additional equation is introduced by seeking the velocity flow field that minimizes
dissipation. It is well known that many problems in physics can be solved using minimums on, for example, travel
distance (light propagation) and energy (free hanging chain). For the interested reader, examples of the application
of the minimum principle in physics is given by Thornton,
6
and a detailed discussion of the fundamental theory of
the calculus of variations is given by Gelfand and Fomin.
7
As a relevant example, the parabolic velocity profile of
fully developed 2-D laminar flow through a circular pipe (i.e. Hagen-Poiseuille flow) can be shown to be a
minimum dissipation solution as shown in Appendix B.
To see how minimum dissipation applies to our case, consider a very small radius ratio such that the return flow
area is large compared to the jet area. The resulting low velocity reverse flow will have negligible effect on the
discharging jet. Thus, initially, the jet should expand like a free jet with constant momentum, i.e. the RHS of Eq.
(26),
3
0 B = . While the momentum is constant under these conditions, the kinetic energy of the jet drops
monotonically (
2
x

) due to the non-zero dissipation. In the free jet case, this process continues ad infinitum
providing for unlimited expansion. In a confined jet, however, the tube constricts the expansion of the jet. The
solution attempts to calculate the value of maximum jet spread using a minimum dissipation approach as described
below.
First, we introduce the dissipation function, u , which is defined by Schlichting,
8
(see Chap. XII Sec. a) as
u
2
0
o
R
u
rdr
r
c | |
|
c
\ .
}
. (32)
It worth mentioning that the dissipation constitutes the RHS of the energy Eq. (10). Next, using the continuity Eq.
(24), a relationship between the forward and backward centerline velocities is formed
$ $
F B
Q
u u f = (33)
American Institute of Aeronautics and Astronautics
8
where
$
( )
$ $
( ) ( )
2
2
2 3 1

1 1
Q
f a a a
| |
+
|
\ .
. Next, from the energy and momentum Eqs. (25) and (26),
respectively, the dissipation, u , and momentum, J , are
$ $ $
$
$
2 2 2
4 5 6
1
F B B b u b u b u

u = + +

(34)
$ $ $ $
( )
$ $ $
( )
2 2 2 2 2
1 2 3
1 1 F B B J c u c u c u = + + . (35)
Finally, using Eq. (33)-(35), the ratio of the dissipation to momentum is expressed by the functional,
$
$
$ $
( )
$ $
( )
2
4 5 6
2 2
2
1 2 3
1
1 1
Q
Q
b f b b
J
c f c c


+ +
u

+ = =
+ +
. (36)
The objective is to find a bounding asymptote for the jet spread,
$
asy = that minimizes this functional, + .
Assuming such a
$
1 asy < is found and substituted into the energy equation (25), and using Eq. (33), one arrives at a
constant gradient of the velocity amplitude when
$ $
asy = ,
$
$ $
( )
$ $
( )
$
$
2 2 3
1 2 3
2
4 5 6
1 1
2 1
3
1
asy asy asy asy
Q
asy
F
oo asy Q
asy asy
Q
asy
asy
b f b b
du
dx f
b f b b

| |
+ +
|
\ .
=
| |
+ +
|
|

\ .
. (37)
It is important to note that the constant gradient of the velocity amplitude implies that the axial centerline velocity
decays linearly along the centerline once
$ $
asy = . Consequently, at a certain distance downstream, after
$
reaches
the value of
$
asy , the flow stagnates.
When
$ $
asy = , the LHS of the momentum Eq. (26) can also be calculated using Eqs. (37) and (33) yielding
$
$
( )
$ $
( )
$ $
( )
$ $
( )
$ $
( )
2 2
2 2
4 5 6 1 2 3
2 2
3
1 2 3
1 1
1
4
3
1 1
asy
asy asy asy asy
Q Q
asy
F
oo Q asy asy asy asy asy
asy Q
b f b b c f c c
u dJ
dx f
b f b b


| |
| |
|
+ + + +
|
|
\ .
|
\ .
=
+ +
. (38)
Equations (37) and (38) describe the flow field in the asymptotic region. The region between the jet discharge and
the asymptotic region is solved next.
As previously discussed, if the radius ratio, c , is much less than 1, the reverse flow near the discharge zone is
negligible and the jet will spread as a free jet in the near field with a zero momentum gradient with an initially linear
jet spread. The jet will continue to expand until it approaches the asymptote. In contrast, if the radius ratio
$
asy c = ,
then the jet boundary should progress along the asymptote from the outset, implying a non-zero momentum gradient
right from the inlet.
A solution that evolves gradually and captures the two scenarios described above is proposed
( )
( )
$ $
( )
$



asy
n
oo
oo
asy oo oo
x dJ dJ
x
dx dx

=
| |
= |
|
\ .
(39)
American Institute of Aeronautics and Astronautics
9
where the positive exponent n is to be determined. It is important to confirm that the proposed solution adequately
describes the two scenarios. At the discharge, 0
oo
x = , the jet width equals the injector radius, thus the jet spread
equals the radius ratio,

c = . If 1 c << , then
( )
$


1
n
oo
asy
x

| |
<< |
|
\ .
, and thus ( )
( )
$ $



asy
oo
oo oo
dJ dJ
x
dx dx

=
<< , which implies
that the jet expands nearly as a free jet with constant momentum. However, as

asy
,
( )
$


1
n
oo
asy
x

| |
|
|
\ .
and
$ $

asy
oo oo
dJ dJ
dx dx
=
as required in the asymptotic region.
Using such a parameterization Eq. (39), the solution will then depend upon an undetermined exponent, n. To
justify this parameterization and find a reasonable value for n, it should be demonstrated that large variations in n do
not significantly affect the salient features of the solution. This will be discussed in the next section.
Notably, the proposed solution overcomes the difficulty associated with not knowing the RHS of the momentum
equation (26), which contains the unknown pressure term. The pressure can be determined retroactively from
3
B ,
defined in Eq. (31), which depends upon the unknown pressure, can be determined as
$ $
3

asy
oo
dJ
B
dx
=
= . (40)
With this method, the pressure can then be calculated independently using
$
$
3
1

1
B
B
oo
dF dp u
B
dx d
q

=
| |
= |
|

\ .
. (41)
The above developed theory can be used to obtain a marching solution with the following steps:
1. Choose arbitrary shape functions,
F
F and
B
F , that satisfy the boundary conditions, Eqs. (12) and (13).
2. From the prescribed flow rate and the shape functions,
F
F and
B
F , calculate the flow parameters (e.g. Re)
that are used in normalizing the variables, see Eqs. (14)-(23).
3. Calculate the a, b and c coefficients using equations (27) through (30).
4. Find

asy
= that minimizes the functional, + , defined by Eq. (36).
5. Start at 0
oo
x = ;

c = ; 1
F
u = ; and calculate
Q
f and
$
B u using Eq. (33).
6. Calculate

oo
asy
dJ
dx
from Eq. (38) using the current values of
F
u and
Q
f .
7. Calculate the elements of A and B in Eq. (31), except for
3
B , which should instead be calculated with Eq.
(40).
8. Integrate Eq. (31) to obtain new values for
oo
x ;

;
F
u ;
$
B u . (Use Eq. (33) to calculate
Q
f ). Integrate Eq.
(41) to obtain the pressure and
9. Go to Step 6, repeating the process until the stagnation region is reached.
Following this solution procedure, the axial dependence of four unknowns (the two velocity amplitudes,
F
u and
B
u ,
the jet spread,

, and the pressure, p ) are determined. These unknowns can then be used to describe the entire flow
field.
American Institute of Aeronautics and Astronautics
10
III. Results and Discussion
The solution employs spatial functions whose boundary conditions were defined (Eqs. (12) and (13)) without
specifying their form. First, a specific form for these functions should be chosen. Since their selection is arbitrary, it
should be demonstrated that the solution is not sensitive to the specifics of the shape functions. To this end, flow
field solutions using two different sets of shape functions will be compared. The two shape functions, named Shape
1 and Shape 2 (for later reference), are given in Eqs. (42) and (43), respectively.
cos
2
1
cos
2
F
B
r
F r
r
F R r
R
t
q q

t q q

| |
= <
|
\ .

| | | |

= + > >
| |


\ . \ .
(42)
( )
( ) ( )
( )
sin
sin 1
1
F
B
r
F r
r
F R r
R
tq
q
tq
t q

q
t q

= <

= > >

(43)
The corresponding coefficients (moments) calculated by Eqs. (27)-
(30) for these functions are given in Table 1. Although the coefficients of
the two profile functions differ, sometimes significantly, the results
obtained with these two velocity function were nearly identical, as will be
shown.
First, the minimum dissipation principle is used to determine the asymptote of the jet spread,

asy
. Figure 4
shows the variation of the dissipation functional + with
$
for the two velocity profile functions, Eqs. (42) and (43).
It is seen that the two profiles show nearly identical behavior for + , suggesting solution independence to the chosen
shape function. The figure indicates that minimum
dissipation is achieved at
$
0.52 = . To keep the
paper within reasonable length, further
investigations will use the Shape 2 profile (Eq. (43))
save a final comparison of the jet penetration
obtained with both shape functions.
As mentioned previously, the exponent, n, in Eq.
(39) is a free parameter, thus, it should be
established that n can assume a large range of values
without significantly affecting the salient features of
the solution. Figures 5 and 6 show the jet spread and
centerline velocity distribution, respectively, along
the tube for various values of n. Figure 5 shows that
for 0,1 n = , the jet spread approaches two different
asymptotes; these asymptotes, however, are not the
one that minimizes the dissipation. For the other
cases ( 1 n > ), though, the solutions predict a spread
with the same asymptote that is proven to minimize
the dissipation. Interestingly, for very large values
of n, the model predicts a free jet expansion along a
straight line that sharply transitions to the
asymptote. Figure 6 shows that the centerline velocity distribution is almost identical for all values of n, specifically
predicting that velocity will decay to zero at 3 8
ii
x ~ . This justifies the use of the ad hoc exponential law in Eq.
(39). For further studies presented in this paper, 2 n = was used.
Shape 1 Shape 2
a
1
0.231335 0.202642
a
2
-0.31831 -0.06456
a
3
-0.63662 -0.13808
b
1
0.109192 0.088275
b
2
-0.21221 -0.00195
b
3
-0.42441 -0.00434
b
4
0.868084 0.719326
b
5
2.467401 0.094307
b
6
4.934802 0.243795
c
1
0.148679 0.123493
c
2
0.25 0.010724
c
3
0.5 0.023558
Table 1. The coefficients defined in
Eqs. (27)-(30) for the two shape
functions, Eqs. (42) and (43).
25
35
45
0 0.2 0.4 0.6 0.8 1
Figure 4. Dependence of the dissipation functional, + ,
upon the jet spread of the flow,
$
, obtained for two
velocity shape functions; A Shape 1, Eq. (42) and Shape
2, Eq. (43).
+

American Institute of Aeronautics and Astronautics


11
Figure 5. Variation of normalized jet spread,
$
, with normalized axial distance
ii
x for radius ratio 0.2 c =
and various exponent values; n = 0 , n = 1 , n = 2 , n = 10 , n = 100 .
Figure 6. Variation of normalized centerline velocity,
F
u , with normalized axial distance,
ii
x , for a radius
ratio 0.2 c = and various exponent values; n = 0 , n = 1 , n = 2 ,n = 10 ,n = 100 .
At this point, the model can be used to predict the effects of geometric and flow parameters on the flow field.
Figures 7 and 8 show the jet expansion and centerline velocity distribution for various radius ratios, c , as a function
of axial distance normalized by the inner radius,
ii
x . Figure 7 shows that the jet spread angle increases with
increasing radius ratio. Assuming, for the sake of discussion, that the radius of the inner tube is unchanged and the
outer tube radius is increased (radius ratio decreases), for small ratios the solution should approximate a free jet
expansion with a jet spread angle inversely proportional to the jet Reynolds number.
8
Thus, as 0 c the jet spread,
, should converge to a single straight line (single spread angle). However, the non-dimensional jet spread,

,
which is normalized by the outer radius (that increases for decreasing c ), converges to straight lines with ever
decreasing slopes, as shown in Fig. 7. Figure 8 shows that the centerline velocity decays faster with increasing
radius ratio. Figures 7 and 8 together show that as the jet expansion nears the asymptote, the centerline velocity
0
0.5
1
0 0.2 0.4 0.6 0.8 1

ii
x
$
u
0
0.5
1
0 0.2 0.4 0.6 0.8 1

ii
x
$

American Institute of Aeronautics and Astronautics


12
decays linearly as discussed previously, see Eq. (37). Notably, Fig. 8 shows that as 0 c , the velocity distribution
approaches the solution of free jet of the form
9
( )
1 1 u x = + .
Figure 7. Variation of normalized jet spread
$
with the normalized axial distance
ii
x for different radius
ratios; c = 0.05 , c = 0.1 , c = 0.2 , c = 0.3 , c = 0.4 , c = 0.5 .
Figure 8. Variation of centerline velocity in the forward direction,
F
u with normalized axial distance
ii
x , for
different radius ratios; c = 0.05 , c = 0.1 , c = 0.2 , c = 0.3 , c = 0.4 , c = 0.5 .
0
0.5
1
0 0.5 1 1.5 2
$
F u

ii
x
0
0.5
1
0 0.5 1 1.5 2
$

$
ii x
American Institute of Aeronautics and Astronautics
13
Figures 9 and 10 show the same data as shown in Figs. 7 and 8 save for the axial scale, which is
io
x instead of

ii
x , see Eq. (20). Significantly, Fig. 9 shows that with such scaling, the jet approaches the asymptote at roughly the
same axial location of about 0.075 for all radius ratios. A similar trend is also apparent in Fig. 10, which shows that
the stagnation point only changes from 0.055 to 0.08 when the radius ratio changes by a factor of 10 from 0.5 c = to
0.05, respectively. This is a most important finding since, thus scaled, all geometries collapse (approximately) to one
significant distance, providing a very powerful measure of a laminar jet discharged into a long dead end tube.
Validation of this finding will be the focus of the next section, where the model predictions will be compared to
CFD simulations.
Figure 9. Variation of the normalized jet spread
$
with normalized axial distance
io
x for different radius
ratios; c = 0.05 , c = 0.1 , c = 0.2 , c = 0.3 , c = 0.4 , c = 0.5 .
Figure 10. Variation of normalized centerline velocity in the forward direction
F
u with normalized axial
distance
io
x , for different radius ratios; c = 0.05 , c = 0.1 , c = 0.2 , c = 0.3 , c = 0.4
, c = 0.5 .
0
0.5
1
0 0.02 0.04 0.06 0.08 0.1

F
u

ii
x
0
0.5
1
0 0.02 0.04 0.06 0.08 0.1
$

$
io x
American Institute of Aeronautics and Astronautics
14
Figure 11 shows the jet penetration as a function of radius ratio scaled with the three different length scales as
defined in Eqs. (19)-(21). It shows
1. the ii-scaled jet penetration varies from infinite as 0 c (not shown, but inferred from a free jet) to
approximately 0.1 when

asy
c = , and
2. the oo-scaled penetration varies from zero as 0 c to approximately 0.028 when

asy
c = .
Thus, the ratio of maximum to minimum penetration for both of these scales is infinite. In contrast, when using the
mixed scaling, io, the penetration distance is relatively constant varying only about plus or minus 15% around the
value of 0.07. This once again illustrates the usefulness of the mixed scaling to represent the behavior of the jet.
Particularly, a rough estimate of the jet penetration would be given as a single value (of 0.07) for any combination of
geometries and inlet velocities.
Figure 11. Variation of three different normalized jet penetration lengths,
, p oo
L ,
, p io
L , and
, p ii
L ,
normalized by Re
o o
R , Re
o i
R , and Re
i i
R , respectively, as a function of radius ratio, c ;
, p oo
L ,
, p oi
L ,
, p ii
L
IV. Model Validation
A. CFD Simulations
The model predictions were compared with results from CFD simulations. The latter were performed using the
commercially available software package, FLUENT 6.2. The problem was simulated as a 2-D axisymmetric laminar
flow, and results were obtained for two Reynolds numbers, Re 50, 100
i
= , and various jet to tube radius ratios,
0.05, 0.1 0.6 c = in steps of 0.1.
Figure 12 shows the geometry of the problem studied. The inlet velocity, at the injector discharge, was assumed
to have a profile of fully developed laminar pipe flow, thus,
2
inlet max 2
1
i
r
u U
R
| |
=
|
\ .
(44)
where U
max
is the centerline velocity and r is the radial distance measured from the axis. The other boundary
conditions were no-slip at the walls (i.e., zero velocity), zero gradients along the axis, and constant ambient pressure
0
0.02
0.04
0.06
0.08
0.1
0 0.1 0.2 0.3 0.4 0.5 0.6
0
2
4
6
8
10
, p oo
L ,
, p oi
L
, p ii
L
c
American Institute of Aeronautics and Astronautics
15
at the outlet. The inlet was offset from the outlet an arbitrary distance, H, to avoid flow reversal (i.e. vortices)
crossing the imposed pressure boundary.
Figure 12. Geometry and boundary conditions used for CFD simulations.
B. Numerical Solution Scheme
FLUENT solves the steady axi-symmetric continuity and Navier-Stokes equations using the finite volume
method with a segregated, implicit, steady-state, laminar, double-precision solver. For pressure-velocity coupling,
the SIMPLE algorithm was used, and for discretization, a first order upwind scheme was used. The grid for the CFD
study comprised a uniform 2D orthogonal mesh, see Fig. 13.
Figure 13. Grid for the radius ratio c = 0.4 with grid spacing / , / 0.067
i i
x R r R A A = .
First, simulations with various mesh spacings were performed to ascertain that the CFD results were not grid
dependent. An example of the results from three different mesh spacings are shown in Table 2, which lists jet
penetration and mass imbalance between the inlet and outlet for each case. It should be noted that the CFD-based jet
penetration differs from the model definition in that it measures the distance for the center line velocity to decay to
5% of the inlet velocity; since, in reality, the velocity
may assume a finite very small (non-zero) value
where it is negligible. The data in suggest that the
results are almost independent of the grid employed
for the normalized grid spacings equal to or smaller
than 0.067 (second and third cases). The results
presented in the following were obtained with a
normalized grid spacing of 0.067.
C. CFD Results
Figure 14 shows the radial distribution of the axial velocity at different cross sections as predicted by the CFD
simulation for a case whose parameters are given in Table 3.

R
o
R
i
L
Wall
Axis Velocity inlet
Pressure outlet
r
H
x
Grid Size
( / , /
i i
x R r R A A )
Jet Penetration
( /
p i
L R )
Mass Imbalance
[ / kg s ]
0.1 8.696 4.86 x10
-12
0.067 8.428 2.83 x10
-12
0.05 8.5 2.20 x10
-12
Table 2. Grid independence study results.
American Institute of Aeronautics and Astronautics
16
As discussed, the boundary condition at 0 x = imposes a parabolic velocity
profile in the region 0
i
r R < < , see Eq. (44). Figure 14 shows that the numerical
scheme did not handle the flow in this region properly. Instead of having a
parabolic profile that reaches zero at / 1
i
r R = , the profile does not reach zero
and, moreover, exhibits a discontinuity. Significantly, this discontinuity skews
the profile, pushing the boundary between the forward and reverse flow outward
to around 0.4, instead of 0.2 as prescribed by the boundary condition. Note that
the discontinuity is absent from subsequent profiles farther downstream.
Consequently, it is expected that the jet spread, measured by the location of the
transition between the forward and reverse flow (i.e. the zero crossing) will be
over-predicted by the CFD simulation.
Figure 14: Velocity profiles at different axial locations plotted versus normalized radial distance, /
o
r R ,
obtained from CFD simulations at various axial locations; 0 x = , 0.05 x = , 0.1 x = , 0.14 x = for a
radius ratio of 0.2 c = .
Figure 15 compares the model and CFD prediction of the jet spread and centerline velocity along the tube for the
same case shown in Fig. 14. The velocities in the figure are normalized by the discharge velocity
max
U from Eq.
(44). The distribution of the jet spread predicted by the CFD simulation clearly shows the abnormality discussed
above; the jet spread is grossly over-predicted at the inlet discharge, 0 x = , followed by an unrealistic narrowing at
the adjacent point. Remarkably, in spite of the discontinuity error in the CFD scheme, it still predicted the
existence of an asymptote at / 0.52
o
r R = , which is identical to the value predicted by the minimum dissipation
method. This confirms the fundamental hypothesis upon which the model was established.
Referring now to the centerline velocity in Fig. 15, the CFD prediction closely follows the general shape
predicted by the model. Significantly, the CFD corroborates the linear velocity decay into the stagnation zone
predicted by the model. It should be noted that while the axial dependence of the centerline and reverse velocities
predicted by both the model and CFD simulation follow similar trends, the predicted jet penetration differs by about
10%. It is quite likely, though, that this discrepancy is the result of the discontinuity error in the CFD simulation.
-0.02
0
0.02
0.04
0.06
0.08
0.1
0.12
0 0.2 0.4 0.6 0.8 1 1.2
u
o
r
R
Parameter Value
o
R 0.07 m
i
R 0.014 m
L 0.5 m
max
U 0.104 m/sec
Re
i
50
c 0.2
Table 3. Simulation
parameters for the data
presented in Figs. 14 and 15.
American Institute of Aeronautics and Astronautics
17
Figure 15. Comparison between predictions by the theory and CFD results for the normalized centerline
forward velocity,
F
u , normalized centerline backward velocity,
B
u and normalized jet spread
$
as a
function of normalized axial distance
ii
x for radius ratio of 0.2;
$
from CFD,
$
from theory;
F
u
from CFD, from theory; from CFD, from theory.
Figure 16. Variation of the normalized jet penetration,
,
Re
p oi p o i
L L R , with radius ratio, c ; Shape 1,
Eq. (42), Shape 2, Eq. (43), and CFD simulations.
, p oi
L
c
0
0.02
0.04
0.06
0.08
0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
-0.2
0
0.2
0.4
0.6
0.8
1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
$
F u ,
$
B u and
$

ii
x
American Institute of Aeronautics and Astronautics
18
Figure 16 compares predictions of the jet penetration obtained by the CFD simulation and model as a function of
the radius ratio. In this comparison, two model predictions are given corresponding to the two shape functions
defined in Eqs. (42) and (43). Figure 16 shows that the use of the two different functions yielded identical jet
penetrations, further demonstrating the robustness of the model.
Figure 16 also shows that the jet penetration predicted by the CFD is in general lesser than that predicted by the
model, including the case where 0.2 c = . This is at odds with the stagnation point location predicted by the CFD
shown in Fig. 15, which shows a longer penetration distance as compared with the model. This is because, as
discussed earlier, the jet penetration calculated from the CFD results was determined by the location where the
centerline velocity reaches 5% of the inlet velocity rather than zero. Taking this into account, the discrepancy
between the model and CFD is even smaller than seen in Fig. 16.
D. Comparison with Experimental Data
Figure 17 shows a comparison of the jet penetration predicted by the model with experimental results obtained
by Eckmann et. al.
4
It should be noted that the normalization procedure used to present the model prediction was
modified to allow proper comparison with the measurements. Figure 17 shows that at low Reynolds number the data
points closely follow the linear trend predicted by the theory. At higher Reynolds numbers, the measured data
indicate lesser penetration. This is expected as the ideal laminar nature of the flow breaks down with increasing
Reynolds number. Noticeably, the trend of the dependence of the penetration upon the radius ratio is confirmed by
the measured data in the entire regime of Reynolds numbers.
0
2
4
6
8
10
12
14
16
18
0 50 100 150 200 250 300 350 400
Figure 17. Comparison between experimental data
4
and present theoretical predictions of the variation of the
jet penetration normalized by the outer tube diameter, /
p o
L D , with Reynolds number, Re
o
, for different
radius ratios; c = 0.1, predicted , measured ^ ; c =0.2, predicted , measured ; c =0.3, predicted
, measured .
V. Conclusion
An analytical model that predicts the behavior of a jet discharged into a dead end tube was developed. At the
core of the model was the hypothesis that the jet expansion reaches an asymptote, and this asymptote could be
deduced using a minimum dissipation approach similar to the principle of least action in mechanics. The model used
assumed velocity profiles whose parameters were determined by continuity, momentum and energy equations. It
was shown that the laminar jet penetration, appropriately normalized, was approximately constant for all geometries
and inlet jet velocities. The predictions of the model were found to be in good agreement with both CFD simulations
and experimental measurements.
Re
o
p
o
L
D
American Institute of Aeronautics and Astronautics
19
Appendix A: Derivation of the Energy Equation, Eq. (10)
Starting with the continuity and axial-momentum equations
1 ( )
0
u rv
x r r
c c
+ =
c c
(A.1)
1 1 u u u p
u v r
x r r r r x
c c c c c | |
+ =
|
c c c c c
\ .
v (A.2)
and multiplying both sides of the momentum equation (A.2) by ur , the left and right hand sides of the expression
become
( ) ( )
2 2 2 2
1 1
2 2
u u u
LHS ru ruv ru rvu u rv
x r x r r
c c c c c
= + = +
c c c c c
(A.3)
2
1 1 1 u p u u p
RHS ur r ur r ur ur
r r r x r r r x
c c c c c c c | | | | | |
= = +
| | |
c c c c c c c
\ . \ . \ .
v v v . (A.4)
Also, from continuity,
( ) rv u
r
r x
c c
=
c c
, (A.5)
which, when substituted into Eq. (A.3), gives
( ) ( )
3 2
1 1
2 2
LHS ru rvu
x r
c c
= +
c c
(A.6)
and thus
( ) ( )
2
3 2
1 1 1
2 2
u u p
ru rvu r ru ur
x r r r r x
c c c c c c | | | |
+ = +
| |
c c c c c c
\ . \ .
v v . (A.7)
Integrating both sides of Eq. (A.7) with respect to r between the limits r = 0 and r = R
o
and noting that the pressure
(thus density) does not vary with r, the following expression is obtained:
( ) ( )
2
3 2
0 0 0 0 0
1 1 1
2 2
o o o o o
R R R R R
u u p
ru dr rvu r dr ru urdr
x r r x
c c c c | | | |
+ = +
| |
c c c c
\ . \ .
} } }
v v . (A.8)
At 0 r = , 0
u
r
c
=
c
and 0 v = , while at
o
r R = , 0 u = , thus
( )
2
0
0
0
o
o
R
R
u
rvu ru
r
c | |
= =
|
c
\ .
. Also, since the density is
constant at every cross section and the net mass flux through the cross section is zero, it follows that
0
0
o
R
urdr =
}
. Eq.
(A.8) thereby reduces to
2
3
0 0
1
2
O o
R R
u
u rdr rdr
x r
c c | |
=
|
c c
\ .
} }
v (A.9)
which gives Eq. (10) used in the theory.
American Institute of Aeronautics and Astronautics
20
Appendix B: Application of Minimum Dissipation to Hagen-Poiseuille Flow
Fully developed, steady, laminar flow in a round tube is considered, so called Hagen-Poiseuille flow. Solving the
steady-state 1-D momentum equation, Eq. (6), for fully developed flow readily provides the classical parabolic
velocity profile (see e.g., Schlichting
8
Ch. I Sec. d). It will be shown that the same profile can be obtained from
minimum dissipation considerations using the fundamental theorem from calculus of variations.
7
Let the velocity ( ) u r satisfy boundary conditions at points 0 r = and
o
r R = . The velocity is subjected to a
constraint, K , of the form
0
( , , ')
o
R
K G r u u dr =
}
(B.1)
where '
du
u
dr
. Consider further the functional, J , whose extremum is to be found, has the form
0
( , , ')
o
R
J F r u u dr =
}
. (B.2)
The theory of calculus of variations implies that a necessary condition for an extremum is the fulfillment of
' '
0
u u u u
d d
F F G G
dr dr

| |
+ =
|
\ .
(B.3)
where
u
F
F
u
c

c
;
'
'
u
F
F
u
c

c
and is a constant, a so-called Lagrange multiplier.
In our case the boundary conditions are
0
' 0
r
u
=
= from symmetry and 0
o
r R
u
=
= from no slip at the wall. The
constraint, K , for the problem is the flow rate, Q, thus,
0
o
R
K urdr Q = =
}
. (B.4)
The functional that represents the dissipation is
2
0
'
R
J u rdr =
}
, which constitutes the RHS of the energy Eq. (A.9),
and is defined in Eq. (32).
Accordingly,
2
'
'
'
0
2 '
0
u
u
u
u
F u r
G u r
F
F u r
G r
G
=
=
=
=
=
=
(B.5)
Substituting Eq. (B.5) into Eq. (B.3) gives
( ) 2 ' 0
d
u r r
dr
+ = (B.6)
which can then be integrated to give a parabolic velocity profile,
2
1
o
r
u c
R
| |
| |
| =
|
|
\ .
\ .
. (B.7)
Eq. (B.7) satisfies the boundary conditions for any constant c. The specific value of c (and thus ) is determined by
the prescribed flow rate Q.
American Institute of Aeronautics and Astronautics
21
References
1
Neumeier, Y., Kenny, J., Weksler, Y., Seitzman, J., Jagoda, J., and Zinn, B. T., Ultra Low Emissions Combustor with Non-
Premixed Reactants Injection, AIAA 2005-3775, 41st AIAA/ASME/SAE/ASEE Joint Propulsion Conference, Tucson, AZ,
2005.
2
Abramovich, G. N., The Theory of Turbulent Jets, translated by Scripta Technica with technical editing by L. H. Schindel,
MIT press, Cambridge, MA, 1963, Chap. 10.
3
Amano, R. S., A Numerical Study of Turbulent Axisymmetric Jets Flowing Into Closed Tubes, ASME J. Energy Resour.
Technol., Vol. 108, No. 4, 1986, pp. 286-291.
4
Eckmann, D., M., Frerichs, T. A., Fogg, N. R. and Lueptow, R., M., Laminar Jet Flow into a Dead-End Tube, ASME
FED, Vol. 237, 1996, pp. 667-672.
5
Eckmann, D. M., and Gavriely, N., Intra-airway CO
2
distribution during airway insufflation in ventilatory failure, J. Appl.
Physiol., Vol. 78, 1995, pp. 546-554 .
6
Thornton, M., Classical Dynamics, 4
th
ed., Harcourt College Publisher, Fort Worth, TX, 1995, Section 7.2.
7
Gelfand, I. M., and Fomin, S. V., Calculus of Variations, translated and edited by R. A. Silverman, Dover, New York, 2000.
8
Schlichting, H., Boundary Layer Theory, 7
th
ed. McGraw-Hill, New York, 1979.
9
Andrade, E. N. da C., and Tsien, L. C., The velocity distribution in a liquid into liquid jet, Proc. Phys. Soc., Vol. 49, No.
4, 1937, pp. 381-391.

You might also like