Anda di halaman 1dari 11

Flow Measurement and Instrumentation 13 (2002) 111

www.elsevier.com/locate/flowmeasinst

Flow conditioners efficiency a comparison based on numerical


approach
Andrea Frattolillo, Nicola Massarotti
Dipartimento di Meccanica, Strutture, Ambiente e Territorio, Universita` degli Studi di Cassino, via G. Di Biasio 43, 03043 Cassino, Italy
Received 4 February 2002; received in revised form 7 March 2002; accepted 28 March 2002

Abstract
Generally, flow conditioner efficiency is experimentally evaluated on the basis of the effects that these devices produce on a
particular type of flowmeter. This approach, beside being very expensive, has certainly obstructed the development of a general
theory and consequently the optimal design of flow conditioners. The present paper presents a comparison of the performance of
different flow conditioners, independently from their effects on particular flowmeters. Therefore, the authors propose the use of
several efficiency parameters to evaluate and compare flatness, uni-directionality and axi-symmetry of the velocity profiles produced
downstream from the conditioners. In order to compare the main conditioners, a brief description of their characteristics is reported.
Then, the numerical analysis is carried out using the finite volume technique. From the results obtained it was possible to calculate
the efficiency of the conditioners under low-level disturbance conditions, such as those produced by an out of plane double elbow
configuration. 2002 Elsevier Science Ltd. All rights reserved.
Keywords: CFD; Finite volumes; Flow conditioners; Installation effects

1. Introduction
Flow measurement is strongly influenced by the velocity profile. Since flowmeters are calibrated and characterised under completely developed flow conditions, perturbations such as swirl, cross-flow, and asymmetry can
produce relevant systematic errors [1].
In practical applications, the fully developed conditions can hardly be obtained, in fact fluid-dynamic perturbations are caused by the elements of the piping itself,
such as elbows, joints and valves. Theoretically, it could
be possible to reduce the influence of such perturbations
using an adequate segment of straight pipe between the
disturbance and the instrument. In practice, due to the
reduced dimensions of the piping, it is useful to use proper flow conditioners [2].
Generally, flow conditioner efficiency is not based on
the velocity profile produced downstream from the
device itself but on the effects produced on a particular

Corresponding author. Tel.: +39-0776-299-668; fax: +39-0776299-393.


E-mail address: massarotti@unicas.it (N. Massarotti).

flowmeter. This approach has certainly obstructed the


development of a general theory and consequently the
optimal design of conditioners [3]. Furthermore, the
number and cost of experimental investigations does not
allow a comprehensive characterisation of flow conditioners, although very interesting laser Doppler investigations were recently carried out [4,5]. The development
of CFD in the last decade [6] has allowed researchers to
use this technique for numerical analysis of installation
effects [7,8] and flow conditioners [9,10].
In the present work the authors carry out a numerical
study on several flow conditioners, independent of their
effects on a certain type of instrument. The analysis is
based on the evaluation of different fluid-dynamic parameters that describe the velocity profile disturbance in a
conduct, with respect to fully developed flow conditions.
An exhaustive analysis is not achievable because of the
variety of both fluid-dynamic perturbations and conditioners. The present work deals with etoile, tube bundle
(AGA), Laws flow conditioners and their performances
under low level perturbation. Obviously, for each conditioner it would be necessary not only to verify the
efficiency with respect to the different types of disturbance and flow conditions, but also to validate the
results experimentally.

0955-5986/02/$ - see front matter 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 9 5 5 - 5 9 8 6 ( 0 2 ) 0 0 0 1 7 - 1

A. Frattolillo, N. Massarotti / Flow Measurement and Instrumentation 13 (2002) 111

Nomenclature
A
D
k
Ka
Kf
Kfm
Kv
Kvm
m

p
r
R
Re
U
Um
U0
Urif
V
W
x
xs
y
ys
z

eKi
m
r

area of the cross section (m2);


pipe diameter (m);
turbulent kinetic energy (m2s2);
asymmetry parameter;
flatness parameter;
momentum of flatness parameter;
swirl angle parameter;
swirl parameter;
mass flow rate (kg/s);
pressure (Pa);
adial coordinate (m);
pipe radius (m);
Reynolds number (Re=rUmD/m);
mean axial velocity (ms1);
average axial velocity (ms1);
maximum axial (inlet section) velocity (ms1);
local velocity for fully developed flow (ms1);
mean tangential velocity (ms1);
mean radial velocity (ms1);
co-ordinate [m];
co-ordinate of the centroid of mass (m);
co-ordinate (m);
coordinate of the centroid of mass (m);
co-ordinate (m);
turb. energy dissipat. rate (m2 s3);
efficiency (6) for the Ki parameters;
dinamic viscosity (Pa s);
fluid density (kg/m3).

This work is a part of a wider research project for the


set up of a procedure for numerical modelling of: (i)
main fluid-dynamic perturbations (e.g. elbows, double
elbows, etc) [11]; (ii) most common flow conditioners
[12]; and (iii) flowmeters most sensitive to installation
effects [13].

2. Characterization of flow conditioners


The use of flow conditioners aims to efficiently reduce
the fluid-dynamic disturbances, caused by the piping,
upstream of flowmeters. Such disturbances may produce
swirl, asymmetry and non-flat velocity profiles. The
efficiency in reaching the mentioned aim cannot be the
only parameter used to compare the different models of
flow conditioners available on the market, because the
achieved effect is a function of the specific disturbance,
as well as the flowmeter used. Furthermore, a correct
evaluation of flow conditioner efficiency should be carried out in a wider technical and economic context, also
taking into account parameters finalised to reduce:

fixed costs of the plant, related to the type of installation and the piping dimensions (length of the
straight conduits upstream and downstream the flow
conditioner)
variable costs, related to pressure losses, maintenance,
and product uncounted by the meter.
In literature, different flow conditioner classifications
exist that generally refer to either the fluid-dynamic performance [14] or the shape of the conditioner [2]. Table
1 presents the main characteristics of the flow conditioners available on the market.
The first classification, based on the functioning
characteristics, distinguishes between turbulent mixing,
whirling action and hybrid behaviour flow conditioners.
Turbulent mixing conditioners remove swirl effects
and asymmetries of velocity profiles through a turbulent
zone that extends up to two diameters downstream from
the conditioner.
In whirling action conditioners the swirl effects and
the velocity profile distortions are removed by the whirl
action due to fluid flow through the grid of the device,

A. Frattolillo, N. Massarotti / Flow Measurement and Instrumentation 13 (2002) 111

Table 1
Main flow conditioner classifications
Conditioners

Description

PLa

ULb

DLc

Plate 0.13 D thick with 35 holes displaced in hexagonal rows


Plate 0.123 D thick with central hole surrounded by smaller holes in two concentric rows.
Num of holes from l:6:12(19)to 1:9:21 (31)
Plate with single central hole surrounded by 4 concentric rows displaced according to the
design
Plate 0.12 D thick with 3 concentric rows with holes equally spaced and diameter lager toward
the external
Plate with single central hole surrounded by an intermediate row with 8 holes and an external
row with 16 holes
3 identical perforated plates, spaced 1 D each other and supported by side arms. The holes
(0.05 D have an hexagonal configuration
2 identical perforated plates, spaced 1 D each other and supported by side arms. The holes
(0.05 D) thicken in the central zone
1 perforated plate with holes having diameters decreasing toward the external

12
13

2
3

9
11

25

23

25

1015

10

22

510

23

Tubes diameter 0.25 D, with intermediate space 1/16 of the transversal section. The tubes 5
with length 10 D are symmetrically displaced
Tubes diameterwith length 24 D. The tubes 9 are designed with an incoming angle (45) to 5
facilitate the flow
The most utilized model comprehends 22 tubes, displaced on concentric rows with smaller
1
diameter toward the external

13

20

22

20

22

2 places, 2 D long, designed in a grid configuration. The combs have a square section with
510
side 0.075 D and 0.45 D depth
Different number of radial vanes divide the conduit in combs. ISO 5167 recommends a length 23
of 2 D

4 little wings and 4 vortex generator elements orthogonally disposed on the internal surface of 1
the conduit and extending till 1/5 D in the transversal section and till 2 D in the axial direction

1 perforated plate connected to an honeycomb section 10 D long


Tube bundles with an internal camera and a radial sequential plate
1 perforated plate obtained combining Etoile vanes stile and circumferential splines attached on
the upstream side
Mounted at the meter inlet and constituted of cylindrical vanes coupled to an upstream plate
with a central hole surrounded by 3 concentric rows

45
23
1

4
7
3

12
7
3

510

Turbulent mixing
Perforated plates
Mitsubishi
Laws
K-Lab
Spearman
Nova
Sprenkle
Kinghorn
New Zanker
Tube bundle
AGA 3
ISO 5167
Stuart
Whirling action
Screens
AMCA
Etoile
Bluff body
Swirl vor tab
Hybrid action
Multi-body
Zanker
Gallagher
Laws vanes
X4X

a
b
c

1
PL: Pressure losses=p / rU2m
2
UL: Upstream length in D
DL: Downstream length in D

whose length (14 diameters), gives rise to non-negligible pressure losses. Compared to the turbulent mixing,
whirling action flow conditioners require a shorter
straight pipe upstream (0.78 diameters), but do not have
the same effectiveness in producing a fully developed
profile.
The hybrid behaviour conditioners are designed to

combine the advantages of both the above mentioned


devices: (i) pressure losses are included from 0.8 to 5
times the kinetic energy of the flow; (ii) straight conduits
required upstream and downstream are, respectively, 37
and 2.512 diameters.
The second classification (second column in Table 1)
distinguishes between the following conditioners:

A. Frattolillo, N. Massarotti / Flow Measurement and Instrumentation 13 (2002) 111

multi-body (catalogued as Type A [2]), represented


by the Zanker model and constituted by a perforated
plate with holes of different diameter (larger in the
middle), followed by the same number of channels
(honeycomb structure). Similar configurations, but
with different number and diameter of the holes
(Laws vaned and X4X models), or with an intermediate chamber between the perforated plate and the
tubes (Gallagher model), presented in Table 1, belong
to this class of flow conditioners
perforated plates (catalogued as Type B [2]) are basically rigid plates with different configurations according to the number and diameter of the holes, disposed
on concentric rows on a single plate (L-Lab, Laws,
Nova, etc.), or to the number of plates (twothree)
always spaced by one diameter and supported by side
bars (Sprenkle and Kinghorn). These conditioners are
characterised by low manufacturing, installation and
maintenance costs, as well as an effective reduction
of swirl effects, but do not have the same effectiveness in reducing the asymmetries of the velocity profile
tube bundle (catalogued as Type C [2]), whose configuration, depending on the number (1922) and
diameters of tubes utilised, is characterised by low
manufacturing and maintenance costs. The reduction
of fluid-dynamic disturbances depend on the tube
arrangement, with a significant and persistent distortion of the velocity profile when installed downstream
from a single elbow or a T joint
screens (catalogued as Type D and E [2]), made of a
grid with square section (AMCA) or with side vanes
(Etoile), require upstream and downstream straight
conduits of, respectively, seven and eight diameters,
while the pressure losses are reduced to 0.8 times the
kinetic energy of the flow
bluff body (swirl-vor-tab) characterised by four fins
orthogonally disposed on the internal surface of the
conduit up to 2 diameters in the axial direction. The
remaining two diameters of the device are characterised by the presence of four vortex generator
elements, with a shaft cone section (30) mounted
with the same orientation of the upstream wings.
Among the new models available, the New Zanker, constituted of a single perforated plate with holes of different diameters, larger in the middle zone, deserves a mention. Recent tests, carried out with water [15], have
shown an effective reduction of the swirl effects and a
sensible decrease of the pressure losses coefficient
(PL=2.8).
The continuous introduction of new conditioners
makes even more complex the choice of designers. This
is due to the lack of a reference standard for conditioner
characterization. For this reason, the achievement of a
widely shared definition of the essential parameters for

the characterization of such devices appears to be fundamental.

3. Flow conditioner efficiency


An efficient flow conditioner should have small pressure losses, reduced axial dimensions and low costs; obviously, it should also be able to reduce fluid-dynamic disturbances, such as swirl, asymmetry and non-flatness.
In order to evaluate the flow conditioner efficiency,
the following absolute and relative efficiency parameters
can be introduced; some of these were already available
in the literature [1621].
3.1. Absolute efficiency parameters
The flatness efficiency parameters (or flatness
number) measure the flatness of the velocity profile, that
is the difference between the effective distorted and the
fully developed velocity profile. Flatness is essential for
instruments that are easily influenced by velocity profile,
such as insertion, turbine and differential pressure
flowmeters.
Two parameters are used: the first, Kf, measures the
difference between the effective non-dimensional flux of
axial momentum and the fully developed one; the
second, Kfm, measures the difference between the effective non-dimensional flux of angular axial momentum
and the fully developed one. The latter is very important
for flowmeters affected by the flatness depending on distance from axis such as turbine flowmeters.
These parameters can be mathematically written as:

Kf

r(U2U2rif)dA

prU2mR2

(1)

r(U2U2rif)rdA

Kfm

prU2mR3

(2)

where A is the conduit cross sectional area, r the density


of the fluid, r the radial coordinate, R the conduct radius,
U, and Urif are the effective and the fully developed axial
velocities respectively, and Um, the average value of the
axial velocity in the section.
The axial vortex efficiency parameters (or axial vortex
numbers) measure the intensity of the axial vortex and
are very important in the flowmeters influenced by swirl,
such as turbines and single-path ultrasonic flowmeters.
Several definitions of swirl angle and swirl number are
used in the literature [1620]: the swirl number is based
on the flux of the angular tangential momentum; the
swirl angle is based on the streamline angle with respect

A. Frattolillo, N. Massarotti / Flow Measurement and Instrumentation 13 (2002) 111

to the pipe axis evaluated at relevant points of the section


[2,21]. In the present work two parameters are used: the
first Kv, defined as the non-dimensional flux of tangential
momentum (basically, such parameter physically represents the tangent to the average swirl angle), the
second Kvm, defined according to Mattingly [4], represents the swirl number:

r|UV|dA

Kv

(3)

prU2mR2

Kvm

r|UV|rdA

(4)

prU2mR3

where V is the tangential velocity.


The asymmetry efficiency parameter measures the
degree of symmetry of the velocity profile and represents
a non-dimensional distance between the centroid of the
mass flow and the axis of the pipe section. This parameter is particularly important for insertion flowmeters
and can be expressed as:
Ka

y2s

2
s

(5)

where xs e ys are the coordinates of the centroid of the


mass flow calculated as:

x dm

xs

3.2. Relative efficiency parameters


On the basis of the absolute efficiency parameters
presented above, it is possible to define a relative
efficiency parameter of a flow conditioner, eki, as the
ratio:
Ki(z)Ki(z)
eKi(z)
Ki(0)Ki(z)

conditioner. In some cases in fact, the conditioners


worsen the velocity profile, as happens, for example, at
its exit, where the axial profile is distorted due to the
conditioner itself.
4. Numerical model
The flow regime for the cases studied in this work is
turbulent. The governing equations are represented by
the conservation of mass and momentum, averaged using
the well known procedure introduced by Reynolds [22].
Since steady state incompressible flow is considered in
this work, the Reynolds averaged NavierStokes
(RANS) equations can be written as:

y dm

ys

Fig. 1. Computational domain: (a) lay-out, (b) computational grid for


the inlet section.

(6)

where Kiand Kirepresent the values of the absolute


parameters calculated for the same pipe configuration
without and with flow conditioner respectively; the coordinate z is evaluated from section z=0 placed immediately downstream the disturbance (Fig. 1a). Therefore,
these parameters measure the efficiency of the conditioners compared to the straight pipe (i.e. without flow
conditioner), as the distance z from the disturbance varies. Obviously, values of the efficiency parameter eki
greater (lower) than zero show an efficiency of the conditioner greater (lower) than the configuration without

mass conservation
U 0

(7)

momentum conservation
1
(UU) p 2mDR
r

(8)

where U=[U, V, W] is the average velocity vector,


D 1 / 2(U UT), and R (uu) represents the
well known turbulent or Reynolds stress tensor, due to
fluctuating velocities u [u,v,w].
One of the main problems in the solution of the RANS
equations is certainly the description of these stresses in
Eq. (8). Several turbulence models for the closure of the
problem have been developed over the last 50 years [23],
and some of them were already tested by the authors in
the numerical simulation of internal turbulent flows [11].

A. Frattolillo, N. Massarotti / Flow Measurement and Instrumentation 13 (2002) 111

The so-called ke model [24] was found to give satisfactory results for the simulation of low level disturbance
and some flow conditioners [9,10].
For all the problems studied, with reference to Fig.
1a, the following boundary conditions are assumed:
walls:
U 0,
V 0, W 0
inlet:

analysis with more degrees of freedom would require


parallel facilities.
5. Results and discussion
The numerical procedure introduced in the previous
section is used to model three types of flow conditioners
(etoile, tube bundle and Laws), all substantially different
in geometry and that belong to the different families of
conditioners presented in Table 1. Fig. 2 presents the

(x y)2 n
(n 1)(2n 1)
1
,
U U0
2n2
D
1

V 0, W 0,
k3/2
3
k [Um0.16(Re)1/8]2, e 0.164
2
0.07D
outlet:

(U,V,W) 0
z
where n is the well-known function of Reynolds number
[22], while and e can be evaluated from the turbulent
intensity and length scale [25].
For the simulation of a viscous layer near the walls,
simplified models are used, the so-called wall functions,
in which a velocity profile is assumed, on the base of
well known empirical relations. The equations for the
turbulent flow are solved outside this layer [25].
The code adopted for the numerical solution of the
problem described above (FLUENT 5.0) is based on the
finite volumes technique [25,26]. Finite volumes allow,
respect to traditional finite difference schemes, modeling
of complex geometry by using unstructured grids. This
is essential for the simulation of geometry, such as that
of the flow conditioner.
The discretized algebraic equations, obtained from the
finite volume procedure, are solved using a semi-implicit
algorithm, SIMPLEC, derived from that originally
devised by Patankar [27]. The second order up-winding
scheme is used for all velocity terms in the momentum
equation, and second order interpolation was also used
for pressure.
The geometry of the flow conditioners is discretized
using unstructured grids. The mesh used for the flow
conditioners is generated using the advancing front type
of procedure; Fig. 1 shows the geometry studied (a) and
the grid used for the inlet section (b). The flow conditioners studied are positioned immediately downstream
from the second elbow, where the origin of the coordinate system is placed (Fig. 1a). About half a million cells
have been used for each of the three computational
domains, but still the mesh sensitivity analysis shows
that a finer mesh would slightly improve the results. An

Fig. 2. Flow conditioner configurations: (a) etoile, (b) tube bundle,


and (c) Laws.

A. Frattolillo, N. Massarotti / Flow Measurement and Instrumentation 13 (2002) 111

different flow conditioners and their placement in the


pipe. The etoile conditioner (Fig. 2a) has four plates equally spaced in the angular direction and a length of two
diameters, as recommended by the ISO 5167; the tube
bundle (Fig. 2b), according to the ISO 5167, is two
diameters long and has 19 tubes; the Laws considered
(Fig. 2c) has 19 holes.
The disturbance configuration under study is a low
level perturbation caused by two elbows on orthogonal
plans (Fig. 1). The second elbow is placed immediately
downstream from the first one, and both have a radius
of curvature 1.5 times the diameter of the pipe. Smooth
walls and a Reynolds number equal to 105 are assumed
throughout the present work and the fluid considered is
water in standard conditions, with a density of 1.0103
kg/m3 and viscosity equal to 1.0103 Pas.
The results obtained from the calculation can be used
to evaluate the efficiency parameters presented in Section 3 for each flow conditioner. In order to evaluate the
relative efficiency parameters the solution for the configuration with no flow conditioner is also needed.
In order to validate the numerical results, the solutions
obtained for single and double elbow have also been
compared, where possible, with experimental data available from literature [4,11,16] or with other numerical
data [9]. Both these configuration are obtained from that
described in Fig. 1, without considering the conditioner
(double elbow) and one of the elbows (single elbow); in
particular the Reynolds number and the other parameters
are the same as the cases with flow conditioners. The
numerical results for both single and double elbow configuration with no flow conditioner have shown excellent
agreement with the experiments [11]. For these cases the
configuration, Reynolds number and the other parameters are the same as the cases with flow conditioners.
From Fig. 3 (single elbow) and Fig. 4 (double elbow) a
comparison of the numerical and experimental velocity
profiles (axial and tangential) is presented for the most
critical section, immediately downstream from the dis-

turbance. In both cases the accuracy of the solution is


better than 5% of the velocity in that section, and
increases further downstream in the pipe [11]. Furthermore, the numerical results for the flow conditioner
simulations show a difference with other numerical solutions [9,16] lower than 1015% in terms of velocity
immediately downstream from the disturbance. However, due to the limited number of studies available in
the literature and the difficulty in retrieving exhaustive
results from documented experiments, it is not possible
at the moment to estimate exactly the uncertainty of the
numerical solution.
The numerical results for the flow conditioners studied
are presented in Figs. 59 in terms of absolute and relative efficiency parameters calculated along the pipe axis.
In particular, each figure shows the efficiency parameters
evaluated for the three flow conditioners considered.
Figs. 5 and 6 show the absolute and relative efficiency
parameters related to the flatness numbers (Fig. 5
presents the absolute and relative parameter for the axial
momentum and Fig. 6 the absolute and relative parameters for the angular axial momentum). In particular,
since flow conditioners generally worsen the profile
flatness for low level perturbations, the parameters Kf
and Kfm are usually very high at the conditioner exit,
and the related efficiencies eKf and eKfm are therefore
lower than zero. From analysis of Figs. 5 and 6 it can
be seen that the Etoile conditioner presents the best performance for low level perturbation, in terms of axial
momentum. However, about 30 diameters downstream
from the disturbance the Laws conditioner presents a
better relative efficiency in terms of angular axial
momentum (Kfm) (Fig. 6b). In any case, the relative
efficiency eKf is greater than zero at least 1020 diameters downstream the conditioner. Therefore, it is generally better to place the meter sufficiently far from the
conditioner when no information is given about the conditioner.
In Figs. 7 and 8 the efficiencies for the axial vortex

Fig. 3. Axial (left) and tangential velocity profiles for the single elbow configuration with no conditioner: comparison with the experimental results
[4] 2.6 diameters downstream from the disturbance.

A. Frattolillo, N. Massarotti / Flow Measurement and Instrumentation 13 (2002) 111

Fig. 4. Axial (left) and tangential velocity profiles for the double elbow configurations (no conditioner): comparison with the experimental results
[4] 2.7 diameters downstream from the disturbance.

Fig. 5. Flatness number Kf (a) and its relative efficiency Kf (b) calculated downstream from the three flow conditioners.

Fig. 6. Flatness number Kfm (a) and its relative efficiency Kfm (b)
calculated downstream from the three flow conditioners.

A. Frattolillo, N. Massarotti / Flow Measurement and Instrumentation 13 (2002) 111

Fig. 7. Axial vortex number Kv (a) and its relative efficiency eKv (b)
calculated downstream from the three flow conditioners.

Fig. 8. Swirl number Kvm (a) and its relative efficiency eKvm (b) calculated downstream from the three flow conditioners.

numbers are presented. Since the swirl parameters Kv


and Kvm decrease rapidly downstream from the flow conditioner, as noticed from Figs. 7a and 8a, their relative
efficiencies eKv and eKvm are about one and four respectively (Figs. 7b and 8b). This means that the flow conditioners are more efficient, from about one to four times,
than the configuration with no conditioner. Furthermore,
the relative efficiencies are of the same order of magnitude for the three flow conditioners considered. Therefore, with respect to swirl, all the conditioners examined
are strongly recommended, as these allow the meter to
be placed immediately downstream from the device with
very low swirl level.
As regards the asymmetry of the velocity profile, all
three conditioners improve the symmetry of the velocity
profile with respect to the configuration without conditioner. However, Fig. 9 shows that the efficiency of the
three conditioners is quite different for the first 25 diameters downstream from the device itself. In fact, while
the tube bundle shows the best relative efficiency for the

asymmetry number, the Laws presents the worst in that


region. Beyond 25 diameters the three conditioners are
equivalent. Therefore, in all situations where a meter is
sensitive to the asymmetry of the velocity profile the use
of a flow conditioner is convenient and can reduce this
effect within 510 diameters.
Finally, we point out that from the present results it
is not possible to quantify the effects of the Reynolds
number on the parameters presented.

6. Conclusions
Numerical methods represent a very useful tool for
the evaluation of flow conditioner efficiency, and can
also be employed in the development of a more general
theory for the evaluation of their efficiency. The work
presented describes a comparison of different types of
flow conditioners under low level perturbation, based on
numerical simulation. Absolute and relative efficiency

10

A. Frattolillo, N. Massarotti / Flow Measurement and Instrumentation 13 (2002) 111

relative efficiency is comparable or lower than the


other two conditioners studied.

Fig. 9. Asymmetry number Ka (a) and its relative efficiency eKa (b)
calculated downstream from the three flow conditioners.

parameters are proposed and calculated on the base of


a numerical method. The results obtained have also been
used to compare the performance of the three conditioners under test (etoile, tube bundle and Laws), as well
as to evaluate their optimal position and use, depending
on the flowmeter sensitivity to the different type of disturbance. In particular, these results show that the conditioners perform differently as regards to flatness, swirl
and asymmetry of the velocity profile produced in
such configuration:
the etoile presents good characteristics in term of
swirl, but the flatness and asymmetry relative
efficiencies are not very high;
the tube bundle shows very good performance in
terms of symmetry, but it also shows a large disturbance of the axial velocity profile;
the Laws is very interesting for the profile produced
far downstream from the double bend, although its

The numerical results obtained macroscopically agree


with the data available from the literature and presented
in Table 1. In particular, as regards the minimum length
to be maintained downstream from the flow conditioner,
the numerical results confirm that the etoile is more
efficient than the other two conditioners considered.
However, a more rigorous approach to the determination
of the conditioner characteristics, and therefore of the
number of diameters to be maintained downstream from
the conditioner, should take into account both the different types of disturbance produced (flatness, swirl and
asymmetry) and the flow meter employed (therefore the
minimum level of flow disturbance acceptable). This
type of analysis, although extremely interesting and
desirable, can only be performed after a proper investigation of the influence of the disturbance parameters on
the main flow meters available.
Although the numerical model has been validated in
the configuration without flow conditioner, a further
validation would be needed in the configuration with the
conditioners, at least for the most critical condition, in
order to evaluate the accuracy of the numerical solutions
and eventually to improve the numerical methodology
developed. However, the accuracy of the numerical solution influences the absolute efficiency parameters to a
greater extent than the relative ones and, therefore, the
comparison carried out can be considered reliable.
At the moment an experimental research is in progress
at the LAMI to investigate the velocity pattern for different pipe configurations [28]. A further improvement of
the numerical model is now under study, such as the use
of different computational grids and new turbulence
models.
Acknowledgements
The authors would like to express their gratitude to
Professor M. DellIsola whose suggestions were fundamental during the preparation of the manuscript.
References
[1] R.W. Miller, Flow Measurement Engineering Handbook, in:
McGraw Hill, New York, 1989.
[2] ISO 5167-1. Measurement of fluid flow by means of pressure
differential devices, ISO Geneve, 1991
[3] J.E. Gallagher, P.J. Lanasa, R.E. Beaty, Development of Gallagher flow conditioner, in: Proceedings of the 7th International
Conference on Flow MeasurementFlomeko 94, Glasgow,
UK, 1994.
[4] G.E. Mattingly, T.T. Yeh, Effect of pipe elbows and tube bundle
on selected types of flowmeters, Flow Measurement and Instrumentation 2 (1991) 413.

A. Frattolillo, N. Massarotti / Flow Measurement and Instrumentation 13 (2002) 111

[5] B. Mickan, G. Wendt, R. Kramer, D. Dopheide, Systematic


investigation of pipe flow and installation effects using laser
doppler anemometryPart II, Flow Measurement and Instrumentation 7 (1997) 151160.
[6] O.C. Zienkiewicz, R. Taylor, The Finite Element Method, in:
Fluid Dynamics, 3, Arnold Publisher, London, 2000.
[7] A. Hilgenstock, R. Ernst, Analysis of installation effects by
means of computational fluid dynamics, Flow Measurement and
Instrumentation 7 (1996) 161171.
[8] M. Holms, J. Stang, J. Delsing, Simulation of flow meter calibration factors for various installation effects, Measurement 15
(1995) 235244.
[9] A. Erdal, A numerical investigation of different parameters that
affect the performance of a flow conditioner, Flow Measurement
and Instrumentation 8 (1997) 93102.
[10] G.L. Morrison, K.R. Hall, J.C. Holste, L. Ihfe, C. Gaharan, R.E.
DeOtte Jr., Flow development downstream of a standard tube
bundle and three different porous plate flow conditioners, Flow
Measurement and Instrumentation 8 (1997) 6176.
[11] M. DellIsola, N. Massarotti, L. Vanoli, Analisi numerica dei disturbi fluidodinamica prodotti allinterno di condotti chiusi circolari, in: Proceedings of the 54th ATI National Conference, Setember 14-17, LAquila, Italy, 1999, pp. 4961.
[12] G. Buonanno, M. DellIsola, A. Frattolillo, N. Massarotti, A. Carotenuto, Numerical analysis of flow conditioner efficiency, in:
Proceedings of the FLOMEKO 2000 Conference, Giugno 4-8,
Salvador, Brasil, 2000.
[13] G. Buonanno, M. DellIsola, N. Massarotti, Numerical evaluation
of distorted velocity profile effects on insertion flowmeters, in:
Proceedings of IMEKO World Conference, September 25-28,
Vienna, 2000.
[14] Flow Conditioners Performance Review, Flow Measurement
Guidance Note n. 11, National Engineering Laboratory, East Kilbride, UK, March 1998.
[15] K. Zanker, D. Goodson, Qualification of a flow conditioning
device according to the new API 14.3 procedure, Flow Measurement and Instrumentation 14 (3) (2000) 7987.

11

[16] B. Mickan, D. Dopheide, Systematic investigations of the


efficiency of flow conditioners, in: Proceedings of the 7th International Conference on Flow MeasurementFlomeko 98,
Giugno 15-17, Lund, Sweden, 1998.
[17] Yeh, T.T., Mattingly, G.E., Laser doppler velocimetry studies of
the pipe flow produced by a generic header, NIST Technical Note
1409, 1995. p. 79.
[18] O. Kitho, Experimental study of turbulent swirling flow in a
straight pipe, Journal of Fluid Mechanics 225 (1991) 445479.
[19] W. Steenbergen, J. Voskamp, The rate of decay of Swirl in turbulent pipe flow, Flow Measurement and Instrumentation 9 (1998)
6778.
[20] J. Haltunnen, Installation effects on ultrasonic and electromagnetic flowmeters: a model based approach, Flow Measurement
and Instrumentation 1 (1990) 287292.
[21] M. Langsholt, D. Thomassen, The computation of turbulent flow
through pipe fittings and the disturbed flow in a downstream
straight pipe, Flow Measurement and Instrumentation 2 (1991)
4555.
[22] H. Schlichting, Boundary-Layer Theory, fourth ed., McGrawHill, New York, 1979.
[23] C.G. Speziale, R.M.C. So, Turbulence modeling and simulation,
in: R.W. Johnson (Ed.), The Handbook of Fluid Dynamics, CRC
Press and Springer-Verlag, USA, 1998 Chapter 14.
[24] B.E. Lauder, D.B. Spalding, Lectures in Mathematical Models of
Turbulence, in: Academic Press Limited, London, 1972.
[25] Fluent 5.0 Users Guide, Fluent Incorporated, Centerra Resource
Park, 10 Cavendish Court, Lebanon, NH 03766, USA, 1997.
[26] H. Versteeg, W. Malalasekera, An introduction to Computational
Fluid Dynamics: the finite volume method, in: Longman Scientific & Technical, England, 1995.
[27] S.V. Patankar, Numerical Heat Transfer and Fluid Flow, in:
Hemisphere/McGraw-Hill, New York, 1980.
[28] G. Buonanno, N. Massarotti, A. Frattolillo, Verifica sperimentale
degli effetti di installazione in condotti chiusi, in: Proceedings
of the 55th ATI National Conference, September 14-17, Matera,
Italy, 2000.

Anda mungkin juga menyukai