You are on page 1of 10

R. P.

BENEDICT

Analytical and Experimental Studies of AS1E


F l u Nozzles

Fellow ASME.

J. S. WYLER
Mem. ASME.
Westinghouse Electric Corp.,
Steam Turbine Division,
Lester, Pa. 19113

In this paper, we first present a state of the art review of the published work concerning
theoretical nozzle discharge coefficients.
Then, we develop a new nozzle discharge
coefficient based on an axisymmetric
boundary layer solution, which in turn is based
on a new axisymmetric potential flow solution. These solutions apply to plenum inlet
installations which offer major advantages over conventional ASME pipe inlet installations as to losses, and as to predictability of the discharge coefficient at the higher Reynolds
numbers encountered in industry. Next, we present new correlations for static pressure
tap errors and apply these to the theoretical (zero lap size) discharge coefficients.
Finally,
we present new experimental data showing how a laminar boundary layer is preserved
and tap error is accordingly minimized for the case of a plenum inlet with
ASME
nozzle contour.

Introduction

include all of the above limitations.

ASMK elliptical nozzles are often used to meter the flow of


water, steam and air (see Fig. 1). Discharge coefficients, based
in part on empirical values, are prescribed by the ASME [1],
and are almost always used to determine actual flow rates.
However, when the discliarge coefficient must be extrapolated
beyond the calibration Reynolds numbers to the higher Reynolds numbers encountered in industry, a more rational basis
is required.
Several theoretical studies of the discharge coefficient have
been published: Shapiro-Smith [2], Ilivas-Shapiro [3], Hall [4],
Cotton-Westcott [5], Leutheuser [6], Soundranayagam [7],
Cotton-Corcich-Sehofield [8], and Benedict-Wyler [9], to name
a few; but they all show severe limitations in their formulations.
For examples: parts of flat plate boundary layer theory have
been used by some, although flow through a nozzle is clearly
axisymmetric with important pressure gradient effects to be
considered; most solutions to date are given in terms of either
all laminar or all turbulent boundary layers, although often a
combination of both prevails; and all solutions concern flow
from inlet plenums, although some of these solutions have been
applied to the more conventional pipe installations. The nowfamous relations of Hall [4], i.e.,
CDL = 1 -

6.92 RD-o-

(1)

0.184 RD-o-s

(2)

This paper begins with the development of a new discharge


coefficient based on a theoretical, stepwise, axisymmetric, laminar-turbulent boundary layer solution of Walz [10]. This is
based on a new axisymmetric potential flow solution obtained
for a plenum inlet installation of the standard ASME nozzle.
We further present new correlations for static pressure tap
errors and apply those to the theoretical (zero t a p size) discharge coefficient to obtain values of Cn one would expect to
measure if throat taps were used. Finally, when possible, our
analytical results are compared with the analytical and experimental results of other published works as well as with new
experimental data t h a t we have recently obtained.

2/3 D

UJJ1 LLIIIIIIIIUIIIIII

Ulllll

and
CDT

Contributed by the Research Committee on Fluid Meters of THE AMERICAN


SOCIETY OF MECHANICAL ENGINEERS and presented at the Winter Annual
Meeting, Atlanta, Ga., November 27-December 2, 1977. Manuscript received
at ASME Headquarters April 27, 1977. Paper No. 77-WA/FM-l.

Journal of Fluids Engineering

-fcFig. 1 Geometry of ASME low p elliptical nozzle

Copyright 1978 by ASME

SEPTEMBER

1978, Vol. 100 / 265

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 05/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Potential Flow Studies


Three primary inputs are required for a boundary layer solution. These are: the position parameter, S/D; the potential
velocity ratio, U/U\; and the geometric parameter, 11/D. These
represent respectively: distance S along the nozzle contour in
terms of a reference length, taken to be the nozzle throat diameter D; the potential velocity U at the outer edge of the
boundary layer in terms of a reference velocity, taken to be
the potential velocity U\ at the nozzle throat in the absence
of a boundary layer; and the radius R of the nozzle boundary,
again with respect to D.
To obtain the initial potential velocity ratio, U/U'i, we at
first used a conducting sheet analog [11]. However, this provided only rough estimates of the information needed since
only two-dimensional flow could be treated in this manner.
Then we considered the use of an electrolytic tank analog [12],
since this could handle the axisymmetric aspects of the problem.
We rejected this approach also because of the complexity of
the tank bottom required. Next, we applied a computerized
numerical relaxation method but abandoned this approach because of the long computer times involved to achieve convergence, and because of the difficulty of modeling the elliptical
surface of the ASME nozzle. Finally, we turned to a powerful,
proprietary, finite element, computer program (WECAN), and
found it entirely satisfactory in providing potential solutions to
the various ASME nozzle installations of interest.
In these potential flow studies, we made use of an analogy,
well known since the times of Fourier, Ohm, and Kirchhoff.
Briefly, for constant density fluid flow (which we are considering), for constant resistivity electric flux (which the conducting
sheet analog offered), and for constant thermal conductivity
heat transfer (which the WECAN approach utilized), the Laplace
equation applies and expresses the variation of field potential.
In two-dimensional cartesian coordinates, for example, this is
simply
*<p f
dx1

<>v

= 0

By1

1
p

dV
dS

~~ p

dT
dS

It is clear from the above three statements t h a t temperature


voltage, and velocity potential are analogous, and in particular
that the required fluid velocity can be obtained either from the
voltage gradient in the electric field, or from the temperature
gradient in the heat transfer field.
Some results of these investigations are given in Figs. 2 and
3. In Fig. 2, we compare the two-dimensional results from the
conducting sheet along with the axisymmetric results from the
finite element program, for a nozzle of /3 = 0.43. From this
work it is clear that we cannot use two-dimensional results for
axisymmetric flows. In Fig. 3, we present axisymmetrie potential flow solutions from the finite element analysis for ASME
low /3 profile nozzles for /3's of 0, 0.43, and 0.6. Also shown in
this figure, for comparison purposes, is the Rivas-fihapiro potential flow solution for this same nozzle. The agreement is
satisfying.

Boundary Layer Discharge Coefficient


The discharge coefficient developed in this section is based
on the Walz [10] stepwise, integral method for approximating
incompressible, axisymmetric boundary layers. Briefly, we require statements for momentum balance, energy balance, velocity profile shape factors, various boundary layer thickness
parameters, and a boundary layer transition criterion.
Momentum Balance,
turn equation:
dd
,_ +
do

We begin with the standard momen-

0(2 + Hn)

d_Z

ds

In fluid fields, the analogous equation is:

kA

dU/dS
TJ
U

pU*

= 0

(3)

where 9 is momentum thickness, Hu is a shape parameter


= S*/9, and 5* is displacement thickness. Equation (3) is
transformed, by Z = 6 RsN and, with the definition r/pU2
= a/R$N, becomes

In electric fields, the governing equation is:


I =

where the potential gradient here is more familiarly the fluid


velocity, i.e., U = dtp/dS.
In heat transfer fields, the analogous equation is:

dUJdS
+ z

- u

{Fl}

Fl

(4)

where the laminar boundary layer is characterized by NL = 1,


the turbulent boundary layer by JVT = 0.268, and all other

dS

NomenclatureA area
CD = theoretical discharge coefficient
(zero t a p size)
COM = measured discharge coefficient
d - pressure tap diameter
1) nozzle throat diameter
e = error in static pressure measurement
H = shape parameter, &**/d
Hu = shape parameter, 8*/0
I = electric current intensity
k = thermal conductivity
m ~ mass flow rate
N characteristic exponent
q = dynamic pressure, rate of heat
transfer
R radius
266 /

Vol. 100, S E P T E M B E R

1978

Rd*
RD
Re
Re*
S

=
=
=
=

tap Reynolds number


throat Reynolds number
momentum Reynolds number
critical value of Ro, defined by (8)
distance along nozzle contour,
measured from leading edge
' T temperature
u = velocity in boundary layer
</ = potential velocity at outer edge
of boundary layer
V = electric potential
V* = friction velocity
x, y =cartesian coordinates
1/ = radial distance from wall
Z = transformation variable
(3 = diameter ratio, Dthroat/^pipo (except in Appendix 1)

5 = boundary layer thickness


5* = displacement thickness
5** = energy thickness
A =
=
V =
p =
T =
<P =

finite difference
momentum thickness
kinematic viscosity
fluid density, electrical resistivity
wall shear stress
potential

Subscripts
1 = inlet
2 = throat
/, = laminar
T = turbulent
Superscripts
i
= ideal or inviscid

Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 05/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

\:i-

0.8

0.6^

2-D RESULTS

j
/

0.4

L^
/

/ /

AXISYMMETRIC
RESULTS
0 = 0.43

n?n
-1.6

~'"H
-1.2

-0.8

Fig. 2

-0.4

S /D

0
S/D

0.4

0.8

1.2

1.6

2.0

Fig, 4 arioys velocity distributions, ASME low 0 nozzle

Potential solutions to low p ASME nozzle

nized when r/q = 0. Transition, by the Walz method, is recognized when log R$ exceeds a critical value defined by
(log Re*) = 2.42 + 24.2 (H -

1.572)

(8)

In the first element that shows transition, the solution is changed


to a turbulent boundary layer by simply redefining Z as ZT
= 6L R$LNT.
From then on, only turbulent parameters are
used.
Iteration.
I t is necessary to modify the initial velocities
( [ / ' x /U'i) obtained from potential flow theory, in accordance
with the displacement thickness obtained from the boundary
layer solution. The velocities are corrected according to the
relation

V,
Fig. 3 Axisymmetric potential solutions to low $ ASME nozzle

undefined quantities are given in Appendix 1. In a stepping


form, for axisymmetric boundaries, (4) becomes

A, +

~zA Hi
Energy Balance.
equation is:
(U

i ++

as

AS

B.Fi

1 +

"ft

R7

In similar fashion, the standard

M
H

dU dS

'

~u~

(F)
(/a)

' - n

(5)

energy

(6)

~z -

where if is another shape parameter = 8**/d, and 8** is energy


thickness. In a stepping form, independent of the axisymmetric
condition, (6) becomes
i-

Bn F- ,

AS

(7)

Method of Solution. Equations (5) and (7) are solved as a


simultaneous system of equations in terms of the thickness
parameter Z and the shape factor H, once S/D, U/U\.
and
R/D are specified for the solid boundaries of interest. A linear
change between these specified points is used to give the required precision, and average values of H and Z are used in
these small AS intervals.
Usually, one starts with a laminar boundary layer by setting
H = 1.572 (the flat plate value), Ar = 1, and Z = 0. One then
proceeds with the stepwise solution until either separation or
transition of this boundary layer occurs. Separation is recog-

Journal of Fluids Engineering

W)

(tv)
D%) /

(9)

\Z> a /_

which is valid in a rigorous sense only in the region where onedimensional flow prevails. This procedure yields a new potential flow velocity distribution from which is obtained a new
boundary layer solution. Such corrections, and such repeated
computer runs are only important at the lower throat Reynolds
numbers (say below 1Q6), as shown in Fig. 4. Even then, the
effect of this iteration is to increase Co by only a slight amount
(on the order of 0.1 percent at l i e = 105).
Obtaining the Discharge Coefficient. Solving the boundary layer equations yields: 8*/D, 6/D, 8**/D, R, and r/q (this latter
being the dimensionless wall shear stress), at every point of
interest along the nozzle contour. There are several formulations available to define the discharge coefficient from this
information. T h a t given by Rivas-Shapiro is:
CD

8*
4 D

+ \D)

y u
8U

J.

(10)

However, most authors (e.g., [4, 6, 7] neglect the bracketed


term in (10) as being negligibly small, and settle for
CD

= 1

5*

(11)

Equations (10) and (11) do not admit t h a t the potential core


velocity must change in the presence of a boundary layer.
When an energy balance as well as a mass balance is applied
across the nozzle, a new formulation for the discharge coefficient results.
A general derivation is given in Appendix 2.
plenum inlet boundary layer solution we use
SEPTEMBER

Thus, for our

1978, Vol. 100 / 267

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 05/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

"-0- S)t

l -

46V/ft

" - 4 52*/ft

4 8 2 **/A!

(12)

11 is clear from (12) that the discharge coefficient will be influenced by the energy thickness (5**) as well as by the displacement thickness (5*).

-"K
^^

Boundary Layer Transition. A boundary layer, which is initially laminar for a plenum inlet nozzle installation, may, at some
location in the nozzle and at some throat Reynolds number,
undergo transition and become turbulent. The boundary layer
solution is, of course, strongly dependent on the choice made
from among the transition criteria, and these include:
1. No transition. That is, the laminar boundary layer may
persist throughout the nozzle at all fiVs. In this regard, Schlichting [13] states: In a region of decreasing pressure (accelerating flow) the boundary layer remains, generally speaking,
laminar."

IUKBULENT B.L. AFTER


SEPARATION BUBBLE

TRANSITION BY SEVERE
LAMINAR SEPARATION
(BUBBLE FORMED)

lO 5
THROAT

lO6
REYNOLDS NO.,

10?

Fig. 5 Discharge coefficient of a plenum inlet ASME nozzle as a


function of transition criterion

3. Transition at the Indifference Point. This earliest of transitions occurs when Re exceeds Rg*, and is identified as the Tollmien-Schlichting indifference point of (8). To his comment
t h a t the boundary layer in accelerating flow is generally laminar, Schlichting adds:
"even a small increase in pressure
causes almost immediate transition." The basic indifference
point is known to move towards higher Ro*'s as the wall roughness and/or free stream turbulence decreases. Walz [10] indicates that Re* increases by as much as 750 at very low turbulence levels. T h e turbulent boundary layer t h a t results from
transition at the indifference point will cause a CDT which
exceeds CDL for Rj> < 10 because such a boundary layer grows
at a smaller rate and is thus thinner than the corresponding
laminar layer at the nozzle throat.
3. Transition Because of Laminar Separation. Whenever r/q
~ 0, the laminar boundary layer can no longer prevail.
In its mildest form, the laminar separation will trigger transition to a turbulent boundary layer in much the same manner as
indicated for the indifference point transition. However, transition by laminar separation usually takes place later than by
indifference. In its most severe form, a separation bubble may
grow and cause a thickening of the resulting turbulent boundary
layer. This has the immediate effect of decreasing CDT with
respect to CDL4. Flat Plate Transition. According to flat plate theory, transition occurs when R = 470. However, this zero pressure
gradient value has little to recommend it for the sharp pressure
gradient flow found in ASME nozzles.
I t is t h e existence of these various transition possibilities
that introduces such large uncertainties in theoretical discharge
coefficients. Fig. 5 describes, by schematic curves, the effect
of the various transition criteria on the discharge coefficient.
One must rely on experimental data to determine which of the
transition criteria applies for a particular nozzle profile.

s /o
Fig. 6 ASME nozzle displacement thickness versus contour position

Boundary Layer Results


Figs. 6, 7, 8, 9, and 10 present our boundary layer results in
terms of *, 6, **, Hu, and r/q for the ASME nozzle.
I n Fig. 6, we compare our results for displacement thickness
with those of Leutheuser [6]. This should be recognized as a
generalized plot in the laminar boundary layer region in the
sense that S*/D \/RD is independent of the throat Reynolds
number. After transition, however, in the turbulent regime,
such as often prevails at the throat tap, no such simple correlation holds.
In Fig, 7, we give a similar generalized plot for momentum
thickness in terms of d/D VRD, and compare it with the results of Rivas-Shapiro [5].
268 /

Vol. 100, S E P T E M B E R

1978

RIVAS-SHAPIRO
LAMINAR

S/ D
Fig. 7 ASME nozzle momentum thickness versus contour position

Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 05/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

AUTHORS'
TURBULENT-2-;^--~~~"~\^ s ^^AUTH0RS' LAMINAR

/"^

^-^'f^

/
RIVAS^ ^ ^ ^ " b
,,-SHAPIRO ^ = ^
/
^.--^
COTTON-WESTCOTT

A//X
/r^

/
/

- H A L L ( F L A T PLATE)

LEUTHEUSER'S DATA

J
/?
THROAT

I06
REYNOLDS NO.,

107

Fig. 11 Various b o u n d a r y layer solutions and an e x p e r i m e n t a l c o m parison for p l e n u m inlet A S M E nozzles


S/ 0
Fig. 8

A S M E nozzle energy thickness versus contour position

^TURBULENT

_FLAT_ PLATE
"TURBULENT

In Fig. 8, the characteristics of energy thickness are given in


terms of 8**/D
VliDFig. 9 shows the shape factor, Hu, from our results compared
with Leutheuser's choice of a constant Hu = 2.554 in the laminar region, and compared with the Oat plate value of Hu = 1.286
in the turbulent region.
In Fig. 10, we present the parameter r/q in terms of t h e
throat Reynolds number. This is a key factor, as will be shown
in the next section, in determining static tap corrections. Given
in this figure are: the turbulent boundary layer values resulting
from the indifference transition; the laminar values resulting
from no transition; and the values obtained by imposing the
lie = 470 transition criterion.
Our primary results are summarized in Fig. 11, where we
present the theoretical discharge coefficient for a plenum inlet
ASME nozzle in terms of the throat Reynolds number. Here,
we compare our two solutions (one based on the laminar separation criterion, and the other based on the indifference point
transition) with the laminar boundary layer solution of RivasShapiro, with the laminar and turbulent flat plate solutions of
Hall, and with the turbulent solution of Cotton-Westcott.
For convenience, we have provided empirical fits for our
solutions and for the Cotton-Westcott solution. These are:
CDL

Fig. 9

= L

5.25

RD~- ;

(13)

Authors

S/ D
A S M E nozzle shape factor versus contour position

j;ood for Reynolds numbers from 105 through 5 X 107,


CDT = 1 ~ 0.130 Ro-o-ass
Authors

(14)

good for Reynolds numbers from 106 through 5 X 10', and


CDT = 1 -

0.106 R D --w

(15)

Cotton-Westcott

I05
THROAT

I06
REYNOLDS NO.,

I07

I0 B

R0

Fig. 10 W a l l shear stress to d y n a m i c pressure p a r a m e t e r for A S M E


throat tap nozzle

Journal of Fluids Engineering

good for Reynolds numbers from 3 X 10 through 5 X 107.


I t is clear from Fig. 11 t h a t significant differences exist between our new boundary layer solutions and previous theoretical
solutions. These differences are to be accounted for mainly
by two factors: the formulation of Co, and the choise of a transition criterion, both of which we have already discussed.
For the plenum inlet ASM E nozzle, only the data of Leutheuser
[6] are available in the literature. This has therefore been
added to Fig. 11. Since these data- were obtained by traversing
the nozzle exit, without the use of throat taps, it can be compared directly with the theoretical curves. It strongly suggests
that a laminar boundary layer (such as described by our equaSEPTEMBER

1978, Vol. 100 / 269

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 05/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

1.00

1
AUTHORS' LAMINAR
THEORY ^-___

_____
~

99

AUTHORS' SIMULATED BUBBLE______.

LLT
UJ

o
o

UJ

5.98X

CO
Q

SOUNDRANAYAGAM
2 " NOZZLE
0 4"NOZZLE

#
.97-

I05
THROAT REYNOLDS NO., Rp

10

Fig. 12 Theoretical and experimental comparisons for a plenum inlet


ISA nozzle

tion (13) exists in the ASME nozzle at Reynolds numbers


below 106.
We have further found, from our boundary layer studies,
t h a t the nozzle profile shape does affect CD. This confirms
what was found by Soundranayagam [7], and means t h a t we
should expect differences between ISA circular arc profiles and
ASME elliptical profiles for example. Fig. 12 presents the
Soundranayagam data compared with our laminar boundary
layer results of (13). Since his data fall far below (13), it
strongly suggests the presence of a laminar separation bubble
in the ISA nozzle. Indeed, reference [7] indicates t h a t such is
t h e case, Soundranayagam having measured and visually observed a separation bubble in this nozzle.
As a matter of interest, we have simulated the separation
bubble effect on CD by imposing on the boundary layer solution
the condition t h a t (d/D) at laminar separation is fixed at an
arbitrary value larger than t h a t just before separation. This is
similar to a suggestion made by Hall [14] in his work on the
orifice coefficient, where a separation bubble is always encountered.
We have also found t h a t the length of the plenum inlet face
of the nozzle does not affect the discharge coefficient. This
confirms what was already found by Leutheuser [6],
As a final observation, we note t h a t the discharge coefficient
based on the boundary layer parameters at the throat tap
location is essentially the same as t h a t based on the boundary
layer at the nozzle exit. This means that we should expect
very little difference between the theoretical (zero t a p size)
discharge coefficients for throat and pipe wall t a p nozzle installations.

Hence, it is important to consider here the effects of such


tap errors on the discharge coefficient, especially so in the high
Reynolds number region where CD must be extrapolated beyond
calibration data.
Rayle [15], has chosen to express the pressure tap size error
(e) in terms of the mean dynamic pressure (q). Others, including: R a y [16], Shaw [17], Jackson [18], Rajaratnam (19),
Rainbird [20], Duffy-Norbury [21], Franklin and Wallace [22],
and Zogg-Thomann [23], have chosen to express the pressure
tap size error in terms of the wall shearing stress ( r ) .
Tap size error is expressed as a function of r since it is believed to arise because of a local disturbance of the boundary
layer. I t is well-documented t h a t near the boundary wall,
there is a region where conditions are determined solely by
r, p, and v (the latter two being the fluid density and kinematic viscosity respectively), and a characteristic length, y.
Dimensional analysis leads at once to a velocity distribution
in the boundary layer given by

= / (A*)

y.

where u is the velocity at any distance y from the wall, and


V* is the friction velocity defined by V r / p , and R* is the
friction Reynolds number defined by V*y/v.
Equation (16) leads to the universal velocity profile, the socalled "law of the wall." I t is universal in the sense that the
same distribution will be obtained in a developing turbulent
boundary layer as in a fully developed turbulent pipe flow.
It follows, from what has been said concerning the universal
velocity profile, that the static pressure size error is also a function of r, p, c, and a characteristic length only. I n tap analysis,
the tap diameter (..) is taken as the characteristic length.
Dimensional analysis leads to two acceptable pairs of nondimensional terms, depending on the grouping of variables chosen. For one grouping, used by Shaw, there results;
-=
r

I t is possible to define a nozzle discharge coefficient in terms


of measurements obtained by traversing across a nozzle throat
according to (12). For examples of such applications, see [1, 6|,
and [7], However, the more usual method is to form an ideal,
incompressible flow rate in terms of a measured pressure drop
across the nozzle (from inlet to throat, or from inlet to a downstream wall t a p ) ; to obtain an actual flow rate by a catchingweighing-timing procedure; and to ratio these two to form the
discharge coefficient. The latter method invo'.ves the use of
static pressure taps.
But, all pressure taps, even if they are constructed perpendicular to the flow direction, are square-edged, are in a region
of essentially zero pressure gradient, have the proper length-todiameter ratio, and do not protrude into the flow region, exhibit
an error in static pressure measurement which depends on t a p
size.

2 7 0 / Vol. 100, SEPTEMBER 1978

(17)

/(R_*)

For another grouping, first proposed by Preston [24], we obtain:


p*> = g ( T * )

(18)

where
R_* =

V*d/v

etf/pv*

= ( - jCR,.*)2

and
=

Static Pressyre Tap Errors

(16)

T<22/>2 =

(R<i*)2

and Rayle provided a third correlation:


= h (d)

(19)

Results of Static Tap Analysis. All the available experimental


data are compared in Figs. 13, 14, 15. Fig. 13 is Rayle's plot
based on (19). I t shows the utter failure of e/q to correlate
static t a p data, even including Rayle's own work. Fig. 14 is
Shaw's plot based on (17). I t shows the very reasonable correlation of static tap data in terms of wall shear stress. However,
note t h a t there is a lack of definition of the faired-in curve at
the higher Reynolds numbers, and this is just where our interest
lies because of the necessity for extrapolating the nozzle discharge coefficient. Fig. 15 is a modified Preston plot based on
(18). There are seen to be two distinct regions of pressure t a p
performance.
One might be considered the laminar region,
and t h e other might be taken as the fully turbulent region.

Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 05/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

&
A,
A
-X
D

.04

.08

.12

HOLE

RAYLE(I949),L/D=4.2
"
, L / D = 15.6
"
,L/D=270
RAYLEII959)
SHAW ( I 9 6 0 )
FRANKLIN a WALLACE (1970)

.16

SIZE, J

.20

.24

.28

(INCHES)
400

Fig. 13 Static tap error as a function of dynamic pressure and tap


size

These log-log plots yield straight line correlations, and in the


fully turbulent region the straight line allows extrapolations to
be made. Rajaratnam made vise of (18) in his work in connection
with static tap pairs, but this was confined to the lower Rd*
region only. H e gave the expression
T*

129 ( p * ) 0.555

800

1200

600

2800

Fig. 14 Static tap error as a function of wail shear stress and tap
Reynolds number

'

7.0AUTHORS ,(21)

which we have expressed in the convenient form


(20)

= 0.000157 (Rd*) i-w*

to describe static tap error in the range, 0 < Rd* < 385. In
the fully turbulent region, we have correlated all the available
data by the new expression.

j g

6.5-

2000 2400

j *

6.0-

lc

5.5-

A
-

(21)

0.269 ( R d * ) o-sss

to describe static tap error, whether such taps are operating in


developing turbulent boundary layers (as found in nozzles) or
in fully developed turbulent pipe flow (as in the usual pipe
wall taps) at tap Reynolds numbers greater than 385.
Note, there is necessarily a transition region between these
two fits where neither is entirely adequate. We suggest t h a t
this region extends from 280 < Rd* < 500.
I t is important to note t h a t Rainbird's data, the only d a t a
t o extend into the higher Reynolds number region, does not
contradict our straight line correlation of (21). I t is further
important to note t h a t Rayle's data correlates far better on
our straight line plot Of Fig. 15 than on his own coordinate
system of Fig. 13. This indicates that e/q is not the valid correlation parameter.
Effect of Static Tap Error on CD. The authors [25] have derived
a relation between a measured discharge coefficient ( C W ) and
a zero tap size theoretical discharge coefficient (CD) which can
be given as
CDM CD +

g^4)'

CD +

CD3

Journal of Fluids Engineering

1/

RAYLE

X SHAW
O RAINBIRO
D

4.0-

FRANKLIN a WALLACE _

3.5-

5.0

5.5
LOG T

Fig. 15 Static tap error as a function of wall shear stress and tap
Reynolds number

Here, CD and (r/q\ are specified by the boundary layer analyses, (13), (14), and Fig. 10 respectively, and the quantity
(e/rh is defined by the correlations of (20) and (21), depending
on the tap Reynolds number,
=

jr/q

'P

\ RB

(24)

(22)

For most practical cases, et/qi, < <C 2 /<J 2 and can be neglectedFor a plenum inlet, (22) reduces to
CDM =

4.5-

Rd*

Co
1 - /3<

5.0-

RAJARATNAM,
(20)
,

(23)

It is significant to note that the static throat tap corrections


which apply to a nozzle operating with a laminar boundary
layer are negligibly small since the wall shear stresses are so
small. However, there will be large and variable effects on
CDM cavised by static throat taps when operating in a turbulent
boundary layer where the wall shear stresses are high.
SEPTEMBER

1978,

Vol.

100 /

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 05/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

271

Table 1 Summary of statistical analysis


1/8" INLET

TAPS

R
1/8" THROAT

ASME

PLENUM
INLET

TAPS

NOZZLE

PERFORATED PLATE
5 0 % SOLIDITY

Schematic of plenum inlet nozzle installation

AUTHORS'
-

THROAT

REYNOLDS

NO.,

a plenum

The /3 Effect
Actually, there are two general problems of interest in applying boundary layer solutions to flow nozzles. The one, plenum
inlet installations, has been dealt with at some length in this
paper. The pipe inlet installation, on the other hand, has been
all but ignored in this presentation, even though it is the most
conventional of applications. This is because no rigorous boundary layer analysis is available at this time. For one thing, the
inlet flow is rotational, and this is counter to t h e boundary
layer concept of an irrotational core flow. For another, and
possibly more important reason, there is a flow separation in
the vicinity of the corner formed between the inlet pipe and the
nozzle face. This separation is to be expected, but it has so
far precluded further boundary layer solution. T h e existence
of this separation strongly suggests the use of a cubic or double
cubic nozzle in place of the elliptical ASME nozzle, as has
often been suggested in the literature, [26, 27, 28, 29].
Several supposed analytical /3 solutions are available in the
literature, [8, 30, 31, 32], but these are in direct opposition to
each other. Although we believe t h a t our CD equation (12A) is
applicable to both pipe and plenum installations, the pertinent
values of S* and 8** must be available before theoretical predictions can bo made. As we have seen, boundary layer solutions
are available for plenum inlets, but we know of no such solutions for (3 flows.

New Experimental Studies


A plenum inlet ASME nozzle installation was designed, built,
and tested to check out some of the ideas presented in the
analytical portion of this paper (see Fig. 16). The plenum
inlet retards the inlet flow, reduces losses across t h e straightening plate, presents essentially irrotational flow t o t h e nozzle,
1978

CI

(%)

1.75

.9888

.0007

4.303

.17

3.01

.9914

4.6

.9920

.0018

2.770

.23

5.2

.9912

.0013

4.303

.32

5.6

.9934

.0016

2.447

.15

6.6

.9955

.0008

4.303

.20

7.0

.9958

.0002

4.303

.05

7.4

.9944

.0005

12.706

.45

7.9

.9946

.0012

12.706

1.08

8.4

.9948

.0006

12.706

.54

8.9

.9940

.0014

2.571

.15

9.6

,9950

and introduces a step at the nozzle inlet edge which initiates


a developing laminar boundary layer along the nozzle profile.
The plenum nozzle installation was tested in a continuous
flow water loop which included a weigh tank system to provide
the actual flow rate. The ideal flow rate was determined via
=

Vol. 100, S E P T E M B E R

N-l,p

DATA

BOUNDARY LAYER
THEORY
THEORY ADJUSTED
FOR STATIC TAP
EFFECTS(d/D=0.06)

Fig. 17 Theoretical and experimental comparisons for


inlet ASME nozzle

272/

fc

(xio-S)

I I N C H ^ 2 . 5 4 CM

Fig. IS

A2

2g pAp Vi
1 - /3*

(25)

where the significant pressure drop across the nozzle was determined from about 25 readings taken during each weighing. A
special 120 inch Statham U-tube manometer, with digital readout to 0.001 inch_mercury under water, was used to provide
Ap. A typical Ap had a standard deviation of 0.023 and a 2
sigma confidence interval of 0 . 0 1 inch Hg under H 2 0 .
Fig. 17 presents our experimental results in terms of CD,
the best value of CD at a given R j . The confidence interval
defined by
CI

(26)

(see for example [33], where the probability p is taken as 95


percent to correspond to the two sigma interval, is summarized
in Table 1.
T h e test points in Fig. 17 show t h a t the initial laminar boundary layer neither undergoes transition nor separates, but remains laminar up to R D = 106. These data, obtained with
water (along with the air data of Leutheuser [6] presented in
Fig. 11) confirm both our new formulation of CD, as defined
by (12), and our theoretical laminar boundary layer solution.
Finally, these data indicate how small is the static tap effect
when operating in a laminar boundary layer.

Summary
1. We have given potential flow velocity distributions for
ASME nozzles for /3's of 0, 0.43, and 0.6, which all show an overacceleration and then a diffusion near the intersection of the
contraction section and the cylindrical throat section of the
nozzle.
2. We have obtained and described in some detail new
boundary layer solutions to an ASME nozzle with a plenum

Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 05/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

inlet. These show the possibility of wide variations in Ci> depending on whether the boundary layer remains laminar, undergoes an indifference point transition to turbulent, or develops
a laminar separation bubble.
3. We have derived a new Co formulation which accounts
for energy thickness in the boundary layer as well as the conventional displacement thickness. Significant differences in Co
are to lie expected when using the more complete analysis as
compared with t h e conventional.
4. We have obtained new static tap error correlations-for
several ranges of t a p Reynolds numbers, and used these in
conjunction with our boundary layer solutions of r/q to predict
these effects on measured discharge coellicients.
5. Our work has been compared favorably with that of
others, where possible, concerning: potential flow velocities,
boundary layer parameters, and static t a p errors.
6. Finally, we have presented results of now experimental
studies of a plenum inlet ASMF. nozzle in water, and these
confirm our theoretical predictions.

Acknowledgments
We would like to express appreciation to our colleagues:
Kddie Phelts, for help in resolving problems encountered in
developing t h e boundary layer computer program; Dr. Steve
Bennett, for providing the WECAN potential solutions; and to
Norm Deming and Les Southall, for suggesting and supporting
this study. Vic Head, Chairman of Fluid Meters Research
Committee SCS-WG1, offered many helpful suggestions during
the review of this paper.

16 Ray, A. K., "On the Effect of Orifice Size on Static Pressure Reading a t Different Reynolds Numbers," Translation in
ARC Rep. T P 498, Nov. 1956. (See also ARC F M 2479).
17 Shaw, R., "The Influence of Hole Dimensions on Static
Pressure Measurements," / . Fluid Mech., Vol. 7, P a r t 4, Apr.
1960, p. 550.
18 Jackson, J. D., "A Note on the Relationship Between
Static Hole Error and Velocity Distribution in the Boundary
Layer," App. Sci. Res., Section A, Vol. I I , 1962, p . 218.
19 Rajaratnain, N., "A Note on the Static Hole Error Problems," J. 'Roy. Aero. Soc, Vol. 70, Feb. 1960, p . 370.
20 Rainbird, W. J., "Errors in Measurement of Mean Static
Pressure of a Moving Fluid Due to Pressure Holes," Quart. Bull.,
Div. Mech. Engrg., Nat, Aero. Est., N R C Rept. D M E / N A E ,
1907 (3).
21 Duffy, J., and Norbury, J. F., "The Measurement of Skin
Friction in Pressure Gradients using a Static Hole Pair," Proc.
Inst. Mech. Enqrg., Vol. 182, Part 3H, 1967-1968, p . 76.
22 Franklin, It. E., and Wallace, J. M., "Absolute Measurement of Static-Hole Error using Flush Tranducers," / . Fluid
Mech., Vol. 42, Part 1, 1970, p . 33.
23 Zogg, H., and Thomann, H., "Errors in Static Pressure
Measurements Due to Protruding Pressure Taps," J. Fluid
Mech., Vol. 54, Part 3, 1972, p. 489.
24 Preston, J. II., "The Determination of Turbulent Skin
Friction by Means of Pitot Tubes," / . Roy. Aero. Soc, Vol. 58,
1954 p. 109.
25 Wyler, J. S., and Benedict, R. P., "Comparisons Between
Throat and Pipe Wall T a p Nozzles," ASME ./. Engrg. Power,
Oct, 1975, p. 569.
26 Rouse, II., and Hassan, M. M., "Cavitation-Free Inlets
and Contractions," Mechanical Engineering, Mar. 1949, p . 213.
27 Redding, T. H., "Flow Characteristics of Metering Nozzles," The Engineer, July 1963, p . 129.
28 -Morel, T., "Comprehensive Design of Axisymmetric Wind
Tunnel Contractions," ASME J. FLUIDS ENQRG., June 1975, p .
225.
29 Hussain, A. K. M. F., and Ramjee, V., "Effects of t h e
Axisymmetric Contraction Shape of Incompressible Turbulent
Flow," A S M E JOURNAL OF F L U I D S ENGINEERING, M a r . 1976, p .

References
1 Fluid Meters - Their Theory and Application, Report of
the ASME Research Committee on Fluid Meters, o'th edition,
1971, Bean, II. S., ed.
2 Shapiro, A. H., and Smith, R. 1)., "Friction Coellicients
in the Length of Smooth, Round Tubes," NACA T N 1785, Nov.
1948
3 Rivas, M. A., and Shapiro, A. II., "On the Theory of
Discharge Coellicients for Rounded-Entrance Flowmeters and
Venturis," TRANS. ASME, Apr. 1956, p . 489.
4 Hall, G. W., "Application of Boundary Layer Theory to
Explain Some Nozzle and Venturi Flow Peculiarities," Proc.
Instn. Mech. Engrs., Vol. 173, No. 36, 1959, p. 837.
5 Cotton, K. C., and Westcott, J. C , "Throat T a p Nozzles
Used for Accurate Flow Measurements," ASME / . of Engrq.
or Power, Oct, 1960, p . 247.
6 Leutheuser, II. J., "Flow Nozzles with Zero Beta R a t i o , "
ASME J. Basic Enqrg., Sept, 1964, p . 538.
7 Soundranayagam, S., "An Investigation into the Performance of Two ISA Metering Nozzles of Finite and Zero Area
Ratio," ASME J. Basic Enqrg., June 1965, p . 525.
8 Cotton, K. C , Carcich, J. A., and Schofield, P., "Experience with Throat-Tap Nozzles for Accurate Flow Measurement,"
ASME J. Engrg. Power, Apr. 1972, p . 133.
9 Benedict, R. P., and Wyler, J. S., "A Generalized Discharge Coefficient for Differential Pressure Type Fluid Meters,"
ASME J. Engrg. Power, Oct. 1974, p . 440.
10 Walz, A., Boundary Layers of Flow and Temperature,
(translated from the German by H. J. Oser), the M.I.T. Press,
1969.
11 Benedict, R. P., "Analog Simulation," Electro-Technology
Science and Engineering Series No. 60, Dec. 1963, p . 73.
12 Benedict, R. P., and Meyer, C. A., "Electrolytic Tank
Analog for Studying Fluid Flow Fields Within Turbomachinery,"
ASME Paper 57-A-120, Dec. 1957.
13 Sehlichting, II., Boundary Layer Theory (English translation by J. Kestin) McGraw-Hill, N. Y., 1955.
14 Hall, G. W., "Analytical Determination of the Discharge
Characteristics of Cvlindrical-Tube Orifices," /. Mech. Engrq.
Science, Vol. 5, No. 1, 1963, p. 91.
15 Itayle, It. E., "An Investigation of the Influence of Orifice
Geometry on Static Pressure Measurement," MS thesis, M.I.T.,
1949. See also, "Influence of Orifice Geometry on Static Pressure
Measurement," ASME Paper 59-A-234, Dec! 1959.

Journal of Fluids Engineering

58.
30 Hall, G. W., Discussion of "An Experimental Investigation of the Flow in a Classical Venturimeter," by D. Lindley,
Proc. Inst. Mech. Engrs., Vol. 184, Part 1, No. 8, 1969-70, p. 147.
31 Au, S. B., "The Prediction of Axisymmetric Turbulent
Boundary Layer in Conical Nozzles," ASME J. App. Mech.,
Mar. 1974, p. 20.
32 "Steam Turbines," A N S I / A S M E P T C 6-1976, An ASME
Performance Test Code, 1976, p . 30.
33 Benedict, It, P., Fundamentals of Temperature, Pressure,
and Flow Measurements, 2nd Edition, Wiley-Interscience, April
1977, p. 191.

APPENDIX

Boundary Layer Relations [10]


AZ

= (Ui/Ui +

iy\

Bz

= (1 - Az (Ui/Ui + 1))/(1 + Fi) (1 - Ui/Ui + 1)

An = (Ui/U + i)F3
Bn

= (1 - Au (Ui/Ui + 1))/(1 + F3) (1 - Ui/Ui + 1)

Fi = 2 + Ar + (1 + Ar) ff

F, = (1 + A) a
F3 = 1 -

Ha

F< = 2/3 Re*-


NNL
NL

NNT

- aH

= 1
= 1
= 0.2317ff - 0.2644 - 87000 (2 - ff)*>

NT

= 0.268

aL

= 1.7261 (H -

aT

= 0.03894 (H - 1.515)0-7

1.51.5) 0 - 7158

PL = 0.1564 + 2.1921 (H - 1.515) 17


SEPTEMBER

1978, Vol. 100 / 2 7 3

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 05/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

(ST = 0.00481 + 0.0822 (H - 1.5)* 81


Hnh = 4.0306 - 4.2845 (II HuT

1.515)-

= 1 + 1-48 (2 - ff) + 104 (2 -

where

fl>7

2 = 0 2?^
.symmetric
"" 1/(1 + A')

i? =

/UK)

Tidr-

Hence, energy flux, applied between inlet and throat, yields!


l/(l + <v)

D \RD

5*

PTTRHP

ir# =

Pl

/'Pi - p A
\

1 - 2

/ pTrIWUjs/2g,

Actual

u)

= KE,

KE,

52*

1 - 2 52*/%

APPENDIX

Equating ideal and actual pressure drop heads yields,U%/U%x.

Discharge Coefficient

Discharge Coefficient

ideally
CD

-.

Continuity, applied between inlet and throat, yields


CD = ( 1 -

(i -

2 -1 - 2 ^ - 2 ^

/?*) ( 1 - 2

s.yii,)

I * ( i - 2 ^ v
i l - 2

?.' = p i i [/i1 = p AtUi

(~-~&^_-

(12a)

Si*/BJ

which, for a plenum inlet, reduces to

where A is geometric area = TTR ,


CD

1 - 2

U' is ideal potential flow velocity,

h*

1 - 2 d2*/Ri
1 - 2 5 2 */2 -

(12)

2 52**/-R

Plenum
and
0 = (fl,/i) =

U\/U\

Energy, under these same conditions, yields


(EV) 2

/Pi - pA
\

/idea!

2(7=

-(1 - 0)

Aetually

With the boundaries displaced by 5*, there results for continuity


m = pA1*U1 = pA2*U, = P7ri?22 ( 1 - 2

d,*/B2)U2

where vi* is area within displaced boundaries = ir(R 5*)2.


U is potential velocity in the core

/32

'1 - 2 W '
i - 2 5i*//ei

r/i

and

o a:

1 _ ._- j ; . . _ d .

DISCUSSION
D, R. Keyset
The authors are to be congratulated on taking yet another
step forward in improving the understanding of the flow coefficients of flow nozzles generally. Especially enlightening is their
inclusion and treatment of the effect of the pressure taps on the
flow measurement. I t has been a recurring weakness in the field
of head class fluid metering to concentrate on the nozzle or orifice
itself, and ignore or circumvent the effects of other parts -of the
measurement system, such as the pressure taps. In this paper

J R-S

I t is necessary to use energy flux, in the actual case, to allow


use of the core velocities, U.
KE

KE

( (KE)dm
(KE)drh
J AA

2(7c

= 27r
f* tihdr
2
9' J 0
^9c

CD

f ->+ r

[//

:- dr
i?

V o l . 100, S E P T E M B E R

1978

\i-2hvru)

The derivation of this equation is given in an ASME paper to be presented at


the 1978 WAM by one of the authors.
3

274 /

If it is in-

'

R-i

/2TrpRm\
\

'Note that the loss term has been neglected in this formulation.
cluded, the resulting discharge coefficient becomes

Naval Ship Engineering Center, Philadelphia, Pa. 19112. Mem. ASME.

Transactions of the ASME

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 05/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like