Вы находитесь на странице: 1из 10

Nuclear Engineering and Design 267 (2014) 3443

Contents lists available at ScienceDirect

Nuclear Engineering and Design


journal homepage: www.elsevier.com/locate/nucengdes

Experimental investigation of ow accelerated corrosion under


two-phase ow conditions
Wael H. Ahmed , Mufatiu M. Bello, Meamer El Nakla,
Abdelsalam Al Sarkhi, Hassan M. Badr
Department of Mechanical Engineering, King Fahd University of Petroleum & Minerals, KFUPM, P.O. Box #874, Dhahran 31261, Saudi Arabia

h i g h l i g h t s

Effect of two-phase ow on ow accelerated corrosion has been investigated experimentally.


Experiments were performed for different orice to pipe diameter ratios.
The effect of ow patterns and mass quality on wear patterns is investigated.
The maximum FAC wear was found at approximately 25 pipe diameters downstream of the orice.
The current study will help FAC engineers to prepare reliable plant inspection scope.

a r t i c l e

i n f o

Article history:
Received 15 May 2013
Received in revised form 30 October 2013
Accepted 2 November 2013

a b s t r a c t
The main objective of this paper is to experimentally study the effect of two-phase ow on owaccelerated corrosion (FAC) downstream an orice. FAC is a major safety and reliability issue affecting
carbon-steel piping in nuclear and fossil power plants. This is because of its pipe wall wearing and thinning effects that could lead to sudden and sometimes catastrophic failures, as well as a huge economic
loss. In the present study, FAC wear of carbon-steel piping was simulated experimentally by circulating
airwater mixtures through hydrocal (CaSO4 1/2H2 O) test sections at liquid supercial Reynolds number, Re = 20,000, and different air mass ow rates. Experiments were performed for a test section with
different orice to pipe diameter ratios (do /D = 0.25, 0.5 and 0.74). The observed ow patterns were compared with the available ow pattern maps. Surface wear patterns downstream the orices were also
analyzed. The maximum FAC wear was found to occur at approximately 25 pipe diameters downstream
of the orice. The obtained results were found to be consistent with those from a single-phase ow study
reported earlier. Moreover, FAC was found to depend on the relative values of the mixture mass quality
and the volumetric void fraction. Lower values of FAC wear rate were obtained for higher values of mass
quality. A modied correlation is developed in order to predict FAC wear rate downstream of the piperestricting orice with an average RMS accuracy of 10%. However, the location of maximum wear rate
is well predicted. The current study is considered as an integrated effort to develop guidelines to FAC
engineers in power plants in order to prepare more reliable plant inspection scope.
2013 Elsevier B.V. All rights reserved.

1. Introduction
Flow accelerated corrosion (FAC) material degradation in carbon
steel piping systems represents one of the major problems in many
industries including nuclear power plants, oil and gas industries,
desalination plants, and many others because of its detrimental
effect on various piping components. It is widely known that the
severe FAC damage normally occurs in tees, elbows, downstream

Corresponding author. Tel.: +966 3 860 7507; fax: +966 3 860 2949.
E-mail addresses: ahmedw@kfupm.edu.sa,
wael.ahmed@gmail.com (W.H. Ahmed).
0029-5493/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.nucengdes.2013.11.073

of control valves, ow elements, reducers or orices. The wall thinning caused by FAC in piping systems may lead to catastrophic
failures of system components and may also result in serious fatalities as reported by Ahmed et al. (2012). Accurate prediction of FAC
rate in a specic application is one of the very complicated problems since it requires detailed investigation of both the soluble iron
production (Fe2+ ) at the oxide/water interface and transfer of the
corrosion products to the bulk ow across the diffusion boundary layer. Consequently, the pipe wall thinning rate due to FAC
depends on a complex interaction of several parameters such as
material composition, water chemistry, and hydrodynamic. Ahmed
et al. (2012) indicated that a signicant research has been conducted on investigating the effect of uid chemical properties on

W.H. Ahmed et al. / Nuclear Engineering and Design 267 (2014) 3443

Nomenclature
A
C
Deff
D
J
MTC
Q
Re
Sc
Sh
Shfd
x
z




pipe cross-sectional area, m2


mass concentration
effective diffusivity, m2 /s
pipe diameter, m
volumetric ux, m/s
mass transfer coefcient, m/s
volume ow rate, m3 /s
Reynolds number, Re = UD/v
Schmidts number, Sc = v/D
Sherwood number, Sh = MTCD/Deff
Sherwood number for fully developed ow
mass quality
axial coordinate, m
volumetric void fraction
kinematic viscosity, m2 /s
density, kg/m3

Subscripts
b
bulk uid value
eff
effective value
wall
w

FAC in nuclear power plants. However, the hydrodynamic effects


of two-phase ows on FAC have not been thoroughly investigated.
Materials degradation due to FAC has been observed under
both single-phase and two-phase ow conditions. In single-phase
water ows, FAC occurs in carbon steel or low carbon-alloy steel
piping/tting at a temperature greater than 95 C, while in the twophase water-vapor ows, FAC also depends on the value of steam
quality (Okada, 2011). Moreover, when the steam quality is low, the
ow-induced corrosion process is referred to as FAC. However, for
high steam quality, the degradation mechanism found to be completely different and referred to as liquid impingement corrosion.
In the later case, the value of ow velocity plays an important role
in the materials damage. For a ow velocity less than 100 m/s, the
damage is mainly due to chemical process, while for an impingement velocity greater than 200 m/s, the degradation is controlled
by a mechanical process. This classication is presented by Okada
(2011) who concluded that liquid droplet impingement velocities
less than 100 m/s found to be incapable of removing the solid corrosion products.
The hydrodynamics parameters controlling FAC in two-phase
ows are expected to be more complex than in the case singlephase ows. This is mainly due to the effect of phase redistribution
and the complex interactions between the gas and the liquid phases
(Kim et al., 2007). In this case, the interaction between the two
phases has a major role in the mass transfer mechanism and consequently the FAC process. Also, the two-phase ow structure (ow
pattern) within the piping component plays an important role in
controlling the FAC process in this component. For example, vapor
or gas bubbles can have a signicant effect on the turbulent kinetic
energy close to the wall and subsequently the mass transfer rate.
This has been supported by the study performed by Jepson (1989)
who showed that high velocity slugs can cause high turbulence and
shear forces at the pipe wall and thus enhance the destruction of
the protective lm which increases FAC rate.
Poulson (1983) experimentally identied four hydrodynamic
parameters directly affect FAC rate. These parameters are the ow
velocity, surface shear stress, turbulence intensity, and mass transfer coefcient. He concluded that mass transfer coefcient is the
most important parameter among other parameters that significantly inuences FAC rate. Furthermore, the experimental work

35

performed by Poulson (1987) for both single-phase and high mass


quality two-phase ows over a wide range of supercial gas and
liquid velocities highlighted two main important conclusions. First,
the modeling using heat transfer analogy in annular two-phase ow
tends to underestimate the rate of mass transfer. Second, the mass
transfer at the bend is at a maximum value at the elbow line of
sight position where the high velocity droplets hitting the bend.
On the other hand, Chexal et al. (1996) reported a correlation
developed based on experimental FAC data for wet steam ow and
showed that the FAC wear rate is signicantly reduced to about 33%
of its original value when the steam quality is increased by only 10%.
This implies that increasing steam dryness is a very effective way
for reducing FAC damage in two-phase ows.
It should be noted that modeling FAC requires an ability to
determine the ow eld and local wall mass transfer rates of corrosion reactants and products. Therefore, the continuity, momentum
and species mass transport equations are required to be solved
using kinetic energy of turbulence-dissipation along with turbulence energy model closures. This has been found to be the most
efcient approach for many industrial applications as reported by
Nesic et al. (1993). In the case of two-phase ow, modeling is
divided into four classes based on the void fraction distributions as
suggested by Tong and Tang (1997). These classes are: (i) bubbles
suspended in the liquid stream; (ii) liquid droplets suspended in the
vapour stream; (iii) vapor and liquid existing intermittently; and
(iv) two immiscible liquids coexisting in a ow. Therefore, different modeling techniques are applied depending on the two-phase
ow patterns.
Remy and Bouchacourt (1992) suggested that the single-phase
ow correlations could be used for two-phase ow by adjusting
Reynolds number with the actual water velocity and taking into
account the void fraction between the steam and water. The wellknown report by Chexal et al. (1996) utilized the idea of Remy and
Bouchacourt (1992) and used Reynolds number dened as:
ReL = VL

dH
L

(1)

where the liquid velocity is expressed as:


VL =

 Q  1x
AL

(2)

The void fraction can be expressed as a function of the two-phase


ow distribution parameters as:
 =

jg 

CO j + Vgj

(3)

where j and jg  are the mixture and vapor volumetric uxes, CO is
the phase distribution coefcient which depends on the two-phase
ow pattern, and Vgj is the vaporgas drift velocity.
Another extensive work carried out on modeling FAC in two
phase ow is developed by Kuo-Tong et al. (1998), who attempted
to predict FAC damage locations on high pressure (HP) turbine
exhaust steam line. The choice of high pressure (HP) turbine
exhaust steam line as a case study was based on the plant measured data of pipe thickness which indicates serious FAC problem
under two phase ow conditions. They proposed a mathematical approach to predict FAC using two-phase ow hydrodynamic
models to investigate the impact of the local parameters on FAC
damage. They reported an improvement of the new approach over
the previous methods used because of the involvement of the
multi-dimensional ow characteristics responsible for FAC wear.
Although the method presented provides more accurate value of
FAC wear rate in two-phase ow, however, it requires intensive
modeling and not reasonable for practical calculation. That is why
empirical-based correlations continue to receive a great attention
due to its simplicity and suitability for use in practical applications.

36

W.H. Ahmed et al. / Nuclear Engineering and Design 267 (2014) 3443

In a similar fashion, Yuh et al. (2008) developed FAC model


for two-phase ow using a case study of extraction piping system
connecting the low-pressure turbine and feed-water heater at boiling water reactors (BWR). They developed a model similar to that
reported by Kuo-Tong et al. (1998) coupled with a corrosion model
in which special treatment of near-wall uid velocity as suggested
by Ferng et al. (1999a,b, 2000) to express FAC wear distributions.
Their FAC results were presented as a function of the local distributions of uid parameters such as two-phase ow velocities, void
fractions, turbulent properties, and local pressure. They concluded
that in two-phase ow condition, the droplet kinetic energy is considered an important parameter that dominates FAC damage. Their
model was qualitatively in good agreement with actual plant data.
Extensive work was performed by Poulson (1999) who reviewed
several complexities in predicting FAC wear rate and he concluded
that a unique relationship between FAC wear rate and mass transfer coefcient exists. He also added to his earlier ndings (Poulson,
1990) that the mass transfer coefcient is the most important
parameter governing FAC wear and the relation between the FAC
wear rate and mass transfer coefcient is not linear as one would
expected. Furthermore, he concluded that FAC wear rate critically
depends on material, environmental and hydrodynamic factors.
In summary, the piping components located downstream of
pipe ttings such as sudden expansion or contractions, orices,
valves, tees and elbows are most susceptible to FAC damage in
both fossil and nuclear power plants systems as explained in the
above literature and in details by Ahmed (2010) and Ahmed et al.
(2012). This is mainly due to the severe changes in ow direction
as well as the development of secondary ow instabilities downstream of these disturbing components, which enhance the mass
transfer rate and consequently FAC. Moreover, FAC in two-phase
ows is strongly affected by the phase redistributions downstream
of these ttings.
Based on the above review, it is clear that the problem of FAC
occurring in two-phase ows downstream of a restricting orice
has not received enough attention despite the aggressive corrosion
in such location. The problem also represents an interesting case
of FAC coupled with two-phase redistribution downstream of the
orice. In this study, the main objective is to evaluate the effect of
two-phase ow on FAC downstream of an orice at different oriceto-pipe diameter ratio. The effect of local ow and mass transfer
parameters on FAC wear rate is also evaluated at liquid Reynolds
number of 20,000. This will help plant engineer during the preparation of plant inspection scope to avoid sudden and catastrophic
failures due to FAC and consequently improve the plant capacity
factor.
1.1. Correlating the mass transfer coefcient
FAC rate is a direct function of the mass ux of ferrous ions
and can be calculated from the convective mass transfer coefcient
(MTC) in the owing water phase. Therefore, FAC rate is expressed
as a function of MTC and the difference between the concentration
of ferrous ions at the oxide/water interface (Cw ) and the concentration of ferrous in the bulk of water phase (Cb ) as:
FAC rate = MTC(Cw Cb )

(4)

As suggested by Poulson (1999), FAC is well correlated with the


mass transfer coefcient and can expressed in terms of Sherwood,
Reynolds and Schmidt numbers as follows:
Sh = a Reb Sc C

(5)

where a, b and c are related to mass transfer which occurs under a


given ow condition and the component geometry. The constant,
a, is geometry-dependent and can be obtained experimentally; b is
usually between 0.3 and 1, and c is typically between 0.33 and 0.4.

Poulson (1999) reported that the actual mass transfer prole


downstream orices had been examined in details by Coney (1980),
who obtained the non-dimensional mass transfer coefcient utilizing the heat transfer data of Krall and Sparrow (1966) and the mass
transfer data of Tagg et al. (1979). The correlation of Coney (1980)
can be expressed as:

ShZ
= 1 + AZ 1 + BZ
Shfd

0.66
ReO

0.0165Re0.86



21

(6)

where Shz is the local Sherwood number at a distance (z) downstream each orice, ReO is the orice Reynolds number, A, B are
empirical constants obtained by Krall and Sparrow (1966) and Tagg
et al. (1979), and Shfd is the fully developed Sherwood number far
downstream the orice i.e. in a straight pipe and reported by Berger
and Hue (1977) in the form:
Shfd = 0.0165Re0.86 Sc 0.33

(7)

2. Experimental facility and test procedure


Experiments are conducted in a ow loop schematically shown
in Fig. 1, which is designed to accommodate different test section
geometries under both single or airwater two-phase ow conditions. Water is supplied from a 100 L tank through a centrifugal
pump driven by a variable speed electric motor. The air is supplied from the main lab compressor and the ow rate is adjusted
by a control valve and measured using an air rotameter with an
accuracy of 2% full scale. The water mass ow rate is controlled
by variable speed pump in addition to a gate valve located on the
water ow line. The water ow rate is measured using a turbine
ow meter with an accuracy of 2% full scale, and the temperature is measured using thermocouples at various locations along
the ow loop. Experiments were performed using a 1-in. diameter
straight tubing at a Reynolds number of 20,000. A straight section
of approximately 75 diameters is installed upstream of the test section to ensure fully developed inlet ow conditions. An additional
straight section of 100 diameters is installed downstream of the
test section.
In the present study, two types of test sections were used to
investigate the effect of two-phase ow on FAC for three orices
with do /D of 0.25, 0.5 and 0.74. The rst set of experiments was
performed, with no mass transfer involved, using acrylic test sections to visualize the two-phase ow redistribution downstream
of the orice at different mass qualities (Fig. 2). The second type
of test section is made of hydrocal (CaSO4 1/2H2 O) downstream of
the orice, as shown in Fig. 3, in order to simulate the FAC wear
pattern in a reasonable test time. This technique has been applied
and tested before by Poulson (1990) and the dissolution of the wall
material depends on the mass transfer of hydrocal from wall into
the bulk ow and used to simulate FAC wear in carbon steel piping
components. Although the changes to the surface occurring from
the mass transfer of the hydrocal to the ow may not be exactly the
same as that would occur in carbon steel piping systems in power
plants, the wear pattern developed is expected to be reasonably
similar to that generated over a longer period of time in carbon
steel piping component as explained by Wilkin et al. (1983). The
dissolution of the hydrocal is considered to depict the dissolution
of the oxide layer in carbon steel pipe. This dissolution leads to
the development of a rough surface that sometimes referred to as
scallops or pours.
In order to determine the saturation limit of the hydrocal
in water, tests were performed using ne particles of hydrocal
and dissolved in the water reservoir and the water conductivity
were recorded as explained by El-Gammal et al. (2012). Running
fully saturated solution through the hydrocal test section showed

W.H. Ahmed et al. / Nuclear Engineering and Design 267 (2014) 3443

37

Fig. 1. Schematic diagram of the experimental facility.

no-wear. This indicates that the measured wear in the present tests
were entirely due to mass transfer and that the hydrocal test sections are not susceptible to mechanical wear. The test section is
prepared by mixing plaster of pairs with deionized water under
vacuum for 2 min to avoid air bubbles and to decrease the porosity
of the nal hydrocal section. The water to plaster ratio was used for
all tests after several trials to obtain reliable test sections.
The overall mass transfer over the entire hydrocal test section is
determined by measuring the electrical conductivity of the circulating water within the ow loop using EU Tech-PC300 meter with
an accuracy of 1%. The operating principle of this technique was
explained in detail by Ahmed et al. (2012). The amount of hydrocal dissolution in the water is obtained through a calibration curve

relating the water conductivity to the amount of dissolved hydrocal. In the present experiments, a maximum concentration of 4%
on volume basis is found to represent the limit where saturation is
reached and the conductivity remains constant. Temperature measurements were used to compensate for the changes in the water
conductivity with temperature. The water conductivity measurements are recorded every 2 min during each experimental run to
determine the overall wear rate within the test section.
The local wear measurements were obtained using FARO-Axis
CMM with Laser Scanner D100 attached to laser power source
of Class 2M. The measured wear is calculated by measuring the
difference between the actual corroded scanned surface and a
CAD model representing the new pipe without corrosion. Wear

Oric
Plaster of Pairs

P1
Pressure Taps

Acrylic Tubing

P2

P3

Note:
Plaster of Pairs Sec on is replaced by acrylic
tubing for ow visualiza on downstream of the
oric

Fig. 2. Test section for two-phase ow experiments.

38

W.H. Ahmed et al. / Nuclear Engineering and Design 267 (2014) 3443

Fig. 3. Hydrocal test section arrangement.

measurements were obtained by scanning the cut test sections with


a measurement accuracy of 037 mm. KUBE software was used to
laser measurement capturing and GEOMAGIC studio software was
used for data processing for each test section. After the cut-test section was scanned, the point cloud data was optimized by reducing
the data noise, over lapping triangular mesh and overhanging data.
Then, the data was merged into polygons and converted into one
stretched water-polygon structure. It should be noted that no data
modication or smoothing operation were carried out in order to
maintain the original data trend. A CAD geometry was created into
the GEOMAGIC studio to represent the new pipe. Then, the wear
proles are identied by the difference between the surface measurements after wear takes place and the original pipe surface. A
sample of scanned proles is shown in Fig. 4.
In order to characterize the pipe surface, scanning electron
microscopy (SEM) analyses were carried out for the casted test
section. The mean diameter of macropores for the prepared test
section found to be within the standard hydrocal (142152 m) as
shown in Fig. 5 and suggested by Villien et al. (2005). The analysis
was carried out on various regions downstream of the orice for
different mass qualities. Specimens were cut from the degraded
test sections, their base and edges were carefully machined and
grinded while preserving the inner surface for analysis. All sections

Fig. 4. Surface wear proles using GEOMAGIC studio.

Fig. 5. SEM of macropores present in the hydrocal test section before testing.

Fig. 6. Inlet ow patterns on Taitel and Duckler map for 1 in. horizontal pipeline.

W.H. Ahmed et al. / Nuclear Engineering and Design 267 (2014) 3443

(a)

0.16

Single phase, water (x = 0)


0.14

Hydrocal Concentration, % by volume, (C - Ci)

39

Air-water two phase (x=0.0011)

d/D = 0.5
Re1ph = 20,000

Air-water two phase (x=0.0021)


0.12

Air-water two phase (x=0.0032)


0.1

0.08

0.06

0.04

0.02

0
0

10

30

20

40

50

60

70

Time, t, (minutes)

Hydrocal Concentration, % by volume, (C - Ci)

(b)

0.14

Single phase water (x=0)

d/D = 0.74
Re1ph = 20,000

Air-water two phase (x = 0.0011)

0.12

Air-water two phase (x = 0.0021)


air-water two phase (x = 0.0032)

0.1

Air-water two phase (x = 0.0032)

0.08

0.06

0.04

0.02

0
0

10

20

30

40

50

60

70

Time, t, (minutes)
Fig. 7. Hydrocal concentration variation with time under both single and two phase ow experiments: (a) orice to pipe diameter ratio of 0.5, (b) orice to pipe diameter
ratio of 0.74.

were coated with gold by ion sputtering to allow better resolution


during SEM analysis. SEM analysis was carried out at a magnication range (10150) to investigate the surface morphology and
the effect of two-phase ow on the shape of wear pattern.

0.012
d/D = 0.25

Wear rate, % by volume/min

0.01
d/D = 0.5
d/D=0.74

0.008

0.006

0.004

0.002

0
0

0.001

0.002

0.003

0.004

0.005

Mass quality, x

Fig. 8. Variation of hydrocal wear rate downstream the orice with inlet mass
quality for all geometries under both single and two phase ow experiments.

3. Results and discussion


For a given inlet water volume ow rate of 0.4 L/s, the air ow
rate is changed from 0.335 to 0.99 L/s. The corresponding supercial velocities of both water and air are calculated and the inlet
two-phase ow pattern upstream of the orice is identied using
the Taitel and Dukler (1976) ow patterns map shown in Fig. 6.
The ow pattern is conrmed with the ow visualization. In the
present test, it was found that all the experiments represent intermittent ow pattern at the orice inlet. The void fraction and mass
qualities of the current test were obtained with the help the pressure and temperature measurements and by using the Chisholm
(1967) equation. At the same time, the void fraction results were
conrmed using multiple capacitance sensors at the inlet section
of the orice.
As explained in detail by Ahmed et al. (2012), a typical ow
behavior through orice is characterized by ow acceleration as
it approaches the orice then separates at the orice sharp edge
forming large vortices downstream. These vortices are responsible
for the reduction in the ow cross-sectional area further downstream up to the minimum value at the vena contracta. Although,
the ow structure in two-phase ow experiences more disturbance
due to the gas deformation, however, the main liquid ow tends

40

W.H. Ahmed et al. / Nuclear Engineering and Design 267 (2014) 3443

Fig. 9. Two-phase ow redistribution doensteam of orifces.

to decelerate toward the ow reattachment point similar to the


single-phase case. Moreover, the velocity at the central core just
downstream of the orice decreases further as the ow develops
downstream. Overall, the maximum centreline velocity increases
within the circulation zone as the orice diameter decreases.
For the orice to pipe diameter ratio of 0.5, wear rates were
measured at different mass quality values ranging from 0.0011 to
0.0032 and the corresponding inlet void fraction varies between
0.377652 and 0.438281. The resulting average wear rates downstream of the orice are shown in Fig. 7 for both single-phase and
two-phase ows. It can be seen that the gradients of the dissolution

Fig. 10. Wear pattern downstream of the orice at different mass qualities
(do /D = 0.5).

rate decrease as the mass quality increases. The gure indicates a


decrease in the dissolution rate by approximately 42% when the
mass quality increases from 0 to 0.0032. This can be attributed to
the decrease in the hydrocal dissolution rate downstream of the orice due to the presence of air where mass transfer in the gaseous
phase is approximately zero.
Similar experiments were performed at the same operating ow
conditions using orice to pipe diameter ratios of 0.25 and 0.74. In
general, it was found that the slope of dissolution rate gradient lines
decrease as the mass quality increases. The corresponding wear
rates downstream the orice found to decrease as the mass quality
increases for all of the three do /D ratios as shown in Fig. 8. This can
be explained on the basis that an increase in the inlet mass quality
has a parallel increase in the void fraction that results in reducing
the mass transfer based on the fact that the mass transfer in the
gaseous phase is insignicant.

Fig. 11. FAC wear rate prole in the hydrocal test section downstream the orice.

W.H. Ahmed et al. / Nuclear Engineering and Design 267 (2014) 3443

41

25
Single phase water (x = 0)

FAC rate (mm/day)

Air-water two phase (x = 0.001)

20

Air-water two phase (x = 0.002)


Air-water two phase (x = 0.003)
Coney (1980) Correlation

15
"Modified Correlation (x=0.001)"
"Modified Correlation (x=0.002)"
"Modified Correlation (x=0.003)"

10

10

12

Z/D
Fig. 12. Variation of overall wear rate downstream the orice with inlet volumetric
void fraction for all geometries under both single and two-phase ow experiments.

As observed by Fossa et al. (2006), the increase in the void fraction causes a signicant decrease in the liquid level (in comparison
with fully developed inlet ow conditions) just downstream the
orice. This consequently led to a reduction in the overall FAC wear
rates in the region downstream the orice. In spite of the increase
in the liquid turbulence downstream of the orice, the mass transfer decreases dramatically because of the increase of the gas phase
contact area with the pipe wall. This has been conrmed using the
ow visualization experiments as shown in Fig. 9. It is clear from
the gure that liquid contact area with the pipe wall becomes less
as the mass quality increases.
For the same mass quality, the liquid ow downstream smaller
orice diameter is subjected to higher deceleration compared to
larger orice diameter, however, the pipe downstream remain to

Fig. 13. Comparison between the experimental data and the correlations for FAC
wear rate downstream the orice (do /D = 0.5).

experience higher wear rate due to the higher liquid turbulence


and consequently higher mass transfer as shown in the cut-test
sections of hydrocal (Fig. 10). The gure shows qualitatively that
the maximum wear occurs in the region Z/D 05 downstream
the orice. Moreover, the wear rate decreases as the mass quality,
x, increases.
The FAC wear rate distributions along the pipe surface downstream the orice plate for ReL = 20,000 are shown in Fig. 11 for
both cases of single-phase and two-phase ow conditions. The gure indicates that the measured FAC wear rate increases steeply
downstream of the orice and reaches a maximum value within
the ow recirculation region (Z/D 05). The rate then decreases
as the ow develops downstream. Moreover, the peak FAC value

Fig. 14. Surface morphology at the location of maximum wear downstream of the orice (do /D = 0.5) at different mass qualities.

42

W.H. Ahmed et al. / Nuclear Engineering and Design 267 (2014) 3443

decrease as the mass quality increases due to the low mass transfer in the gaseous phase. For single phase ow, the peak value
slightly decreases by about 6% and 42% for x = 0.002 and x = 0.003,
respectively. The location of the maximum FAC wear rate remains
unchanged.
The increase in local liquid velocity has been reported to signicantly affect FAC wear rate downstream the orice under
single-phase ow condition (Chexal et al., 1996). In the present
study, the effect of increasing the void fraction downstream the
orice outweighs the corresponding effect of increasing the local
Reynolds number of the liquid phase and the overall wear rate
decreases as shown in Fig. 12. Hence, the FAC wear rate downstream
the orice is reduced for the case of two-phase ow in comparison
with the corresponding single-phase ow condition. However, the
location of the maximum wear remains the same due to the effect
of higher turbulence in this region.
The data presented in Fig. 12 are used to correlate the effect
of two-phase mass quality on the non-dimensional mass transfer
coefcient previously correlated by Coney (1980). A modied correlation of existing Coney (1980) correlation is proposed to predict
the mass transfer coefcient in the form:

ShZ
= 1 + AZ 1 + BZ
Shfd

Re0.66
O
0.0165Re0.86



21

f (x)

In general, FAC wear rate downstream the orice was found


to decrease under two-phase ow conditions compared to the
single-phase ow at the same inlet liquid Reynolds number.
The location of maximum wear was found to remain within 5
diameters downstream the orice for both single and two-phase
ows.
The mass transfer and consequently the FAC wear rate downstream the orice decreases as the inlet mass quality of increases.
Acknowledgements
The authors would like to thank King Abdulaziz City for Science
and Technology (KACST) for funding this work under the National
Science Technology Plan (NSTIP) grant no. 11-ADV1619-04. Also,
the support provided by the Deanship of Scientic Research (DSR)
at King Fahd University of Petroleum & Minerals (KFUPM) is gratefully acknowledged. The authors thank Mr. Malik for performing
SME tests, and to Mr. Ahmed Abdel Rehim for fabricating the hydrocal test sections.

(8)
References

where f (x) is a function of mass quality and obtained by performing regression analysis of the data presented in Fig. 10 for the
same diameter ratio (do /D) with a correlation coefcient R = 0.87. It
should be noted that the mass quality multiplier expressed in Eq. (9)
is valid for 0 < x < 0.1. As shown in Fig. 13, the modied correlation
is used to calculate the FAC wear rate from the non-dimensional
mass coefcient downstream of the orice as expressed in Eq. (8)
along with the mass quality multiplier expressed as:
f (x) = (1 x0.05 )

single phase and two-phase ow experiments, the following can be


concluded:

(9)

The value of species concentration cw along the wall (equal to


0.275 g/100 g as specied in the hydrocal manufacturers properties tables), and the species concentration in the bulk liquid ow
beyond the diffusive boundary layer, cb is calculated as explained in
details by Ahmed et al. (2012). The modied correlation was found
to correlate the local FAC wear rate within 10% based on the average Root Mean Square (RMS) accuracy. Furthermore, the location of
maximum wear rate is well predicated by the modied correlation.
This is considered an acceptable accuracy since the main objective
is to identify the location of maximum wear downstream of the
orice for inspection and mentoring purposes.
Also, it should be noted that the surface morphology obtained
using SME for the degraded surface under both single- and twophase ows helps in explaining the mechanism of mass transfer at
the pipe wall. Fig. 14 shows the surface morphology at the location
of maximum wear downstream of the orice for different mass
qualities. In can be clearly seen that the surface has pits initiated
mainly from air bubbles exist in the test section hydrocal material.
These pits were reduced in the case of two-phase ow where mass
transfer rate is reduced. As the mass quality decreases to the singlephase liquid case (Fig. 14a), the pours evolve and increase in size
and the surface roughness dramatically increased. The size of these
pours also decreases as the mass quality increases.
4. Conclusion
The effect of two-phase ow on wall mass transfer rate downstream orices was investigated experimentally using hydrocal test
sections. Different orice to pipe diameter ratios were considered
at supercial water Reynolds number of Re = 20,000. Comparing the

Ahmed, W.H., 2010. Evaluation of the proximity effect on ow accelerated corrosion.


Annals of Nuclear Energy 37, 598605.
Ahmed, W.H., Bello, M.M., Al-Sarkhi, A., El Nakla, M., 2012. Flow and mass transfer
downstream of an orice under ow accelerated corrosion conditions. Nuclear
Engineering and Design 252, 5267.
Berger, F.P., Hue, K.F.F.L., 1977. Mass transfer in turbulent pipe ow measured by
the electrochemical method. International Journal of Heat and Mass Transfer
20, p1185.
El-Gammal, M., Ahmed, W.H., Ching, C.Y., 2012. Investigation of wall mass transfer
characteristics downstream of an orice. Nuclear Engineering and Design 242
(6), 353360.
Chisholm, D., 1967. A theoretical basis for the LockhartMartinelli correlation for
two-phase ow. International Journal of Heat and Mass Transfer 10, 17671778.
Chexal, B., Horowitz, J., Jones, R., Dooley, B., Wood, C., Bouchacourt, M., Remy, F.,
Nordmann, F., Paul, St.P., 1996. Flow-Accelerated Corrosion in Power Plants,
Electric Power Research Institute Report No. TR-106611.
Coney, M., 1980. CERL Internal Report, Ref: RD/L/N197/80.
Ferng, Y.M., Ma, Y.P., Ma, K.T., Chung, N.M., 1999a. A physical model to predict wear
sites engendered by ow-assisted corrosion. Nuclear Technology 126, 319330.
Ferng, Y.M., Ma, Y.P., Ma, K.T., Chung, N.M., 1999b. A new approach to investigate
erosion/corrosion phenomenon through the local ow models. Corrosion 55 (4),
332342.
Ferng, Y.M., Ma, Y.P., Chung, N.M., 2000. Application of local ow models in predicting distributions of erosioncorrosion locations. Corrosion 56 (2), 116.
Fossa, M., Guglielmini, G., Marchitto, A., 2006. Two-phase ow structure close to
orice contractions during horizontal intermittent ows. International Communications in Heat and Mass Transfer 33, 698708.
Jepson, W.P., 1989. Modelling the transition to slug ow in horizontal conduit. Canadian Journal of Chemical Engineering 67, 731740.
Kim, S., Park, J.H., Kojasoy, G., Kelly, J.M., Marshall, S.O., 2007. Geometric effects of 90degree elbow in the development of interfacial structures in horizontal bubbly
ow. Nuclear Engineering and Design 237, 21052113.
Krall, K.M., Sparrow, E.M.J., 1966. Heat transfer. Transactions of the ASME, Journal
of Heat Transfer.
Kuo-Tong, M.A., Ferng, Y.M., May, P., 1998. Numerically investigating the inuence of
local ow behaviours on ow-accelerated corrosion using two-uid equations.
Nuclear Technology 123, 90102.
Nesic, S., Adamopoulos, G., Postlethwaite, J., Bergstrom, D.J., 1993. Modelling of turbulent ow and mass transfer with wall function and low-Reynolds number
closures. Canadian Journal of Chemical Engineering 71, 2834.
Okada, H., 2011. Evaluation methods for corrosion damage of components in cooling systems of nuclear power plants by coupling analysis of corrosion and ow
dynamics (V): ow-accelerated corrosion under single- and two-phase ow
conditions. Nuclear Science and Technology 48 (1), 6575.
Poulson, B., 1999. Complexities in predicting erosion corrosion. Wear 233235,
497504.
Poulson, B., 1983. Electrochemical measurements in owing solutions. Corrosion
Science 23 (4), 391430.
Poulson, B., 1987. In: Strutt, J.E., Nicholls, J.R. (Eds.), Predicting the Occurrence of
Erosioncorrosion, Plant Corrosion: Prediction of Material Performance. Ellis
Horwood Ltd., Chichester, England.
Poulson, B., 1990. Mass transfer from rough surfaces. Corrosion Science 30 (6/7),
743746.

W.H. Ahmed et al. / Nuclear Engineering and Design 267 (2014) 3443
Remy, F.N., Bouchacourt, M., 1992. Flow-assisted corrosion: a method to avoid damage. Nuclear Engineering and Design 133, 2330.
Tagg, D.J., Patrick, M.A., Wragg, A.A., 1979. Heat and mass transfer downstream of
abrupt nozzle expansions in turbulent ow. Transactions of the Institution of
Chemical Engineers 57 (12), 176181.
Taitel, Y., Dukler, A.E., 1976. A model for predicting ow regime transitions in horizontal and near-horizontal gasliquid ow. AIChE Journal 22, 47.
Tong, L.S., Tang, Y.S., 1997. Boiling Heat Transfer and Two-Phase Flow, Series in
Chemical and Mechanical Engineering. Taylor & Francis Group, Washington D.C.,
USA.

43

Villien, B., Zheng, Y., Lister, D.H., 2005. Surface dissolution and the development of
scallops. Chemical Engineering Communication 192, 125136.
Wilkin, S.J., Oates, H.S., Coney, M.W.E., 1983. Mass transfer in straight pipes and 90
deg bends measured by the dissolution of plaster. Central Electricity Generating
Board, Technology Planning and Research Division, Central Electricity Research
Laboratories, TPRD/L/2479/N83.
Yuh, M.F., Yung, S.T., Bau, S.P., Long, W.S., 2008. A two-phase methodology to predict
FAC wear sites in the piping system of a BWR. Nuclear Engineering and Design
238, 21892196.

Вам также может понравиться