Sei sulla pagina 1di 7

H618

Journal of The Electrochemical Society, 163 (7) H618-H624 (2016)


0013-4651/2016/163(7)/H618/7/$33.00 The Electrochemical Society

Solid-Liquid Mass Transfer under Flow Boiling Condition


Chen Shen,a Artin Afacan,a Jing-Li Luo,a,,z Jaganathan Ulaganathan,b and Stan J. Klimasb
a Department of Chemical and Materials
b Canadian Nuclear Laboratories, Chalk

Engineering, University of Alberta, Edmonton, AB T6G 2V4, Canada


River, Ontario K0J 1J0, Canada

The study of mass transfer behavior under a flow boiling condition has not been reported due to the complexities of the phenomenon
and experimental difficulties. In the present study, a novel high temperature and pressure experimental flow loop as well as measuring
techniques were developed to determine the mass transfer coefficient under the flow boiling condition to a level that no one else has
achieved according to open literature. A new mass transfer correlation was proposed and proven to be suitable for representing the
flow boiling mass transfer behavior. This work will provide a good base for further flow boiling mass transfer studies.
2016 The Electrochemical Society. [DOI: 10.1149/2.0081608jes] All rights reserved.
Manuscript submitted January 8, 2016; revised manuscript received May 2, 2016. Published May 12, 2016.

Flow boiling is a complex condition involving elevated pressures


and temperatures as well as vapor generation and vapor-liquid twophase flows.13 This condition is commonly encountered in various
industries such as nuclear, energy, heating and cooling processes.4,5
Mass transfer under the flow boiling condition is a fundamental issue
that can cause many problems, such as fouling in the boilers;6 and the
material degradation of the wire plating cells,7 the electrochemical
reactors,8 as well the heat exchange tubing in the steam generators.9
The electrochemical approach is a widely used method to investigate mass transfer behavior under various conditions.10,11 The mass
transfer behavior in annular shaped tubing, which has the same geometry as in the present work, under isothermal conditions has been reported previously.12 Ross and Wragg13,14 investigated free and forced
convective mass transfer around the horizontal electrode and annuli.
The mass transfer rates were determined using the electrochemical
method for the copper deposition reaction from an acidified copper
sulfate solution. Ross and Wragg13,14 proposed a semi-empirical equation for predicting the mass transfer rate on the inner wall of an annular
flow cell for both laminar and turbulent flow regions.
However, the mass transfer behavior under flow boiling condition studies is very limited in the open literature due to experimental
difficulties. Wragg and Nasiruddin15 attempted to measure the mass
transfer rate of copper ions under a subcooled boiling condition. They
stated that the limiting current value could not be measured under a
boiling condition due to the intense local turbulence caused by the bubbles formed during boiling. Although the flow boiling phenomenon
has been extensively studied in the field of heat transfer,1618 the traditional analysis between heat and mass transfer cannot be applied to
the flow boiling condition due to the boundary layer discontinuity.19
Also, the existing mass transfer correlation for a gas evolving electrode
cannot be applied to a boiling electrode because the energy transfer
on a boiling surface is different from the gas generation electrode.20
Thus there is a clear knowledge gap for the mass transfer behavior
under flow boiling conditions both experimentally and theoretically.
The reason for the lack of investigation is the experimental difficulties
of conducting measurements on the flow boiling surface.
One of the major problems with the non-isothermal electrochemistry studies is the heating method for the working electrode. Several
different heating methods have been reported to heat the working electrode in the literatures. Beckmann et al.21,22 used direct AC heating
wire electrodes to analyze the temperature-dependence of electrochemical reactions. Their results appear to be specific to the natural
convection at the microelectrode. Since they used a Pt wire with a
diameter of 25 m as an electrode, the convection and diffusion behavior may not be applicable to at larger scale (diameters more than 10
cm). Marken et al.23 used microwave radiation heating of the working
electrode to detect anodic stripping of cadmium. This was a promising
technique, however they did not conduct their measurements under
boiling and activation conditions. Sarac and Wragg24 immersed a preheated nickel rod in the ferri-/ferrocyanide NaOH solution, to measure
Electrochemical Society Member.
z
E-mail: jingli.luo@ualberta.ca

the mass transfer behavior in a short time (10 s) at the pool boiling
surface. Their result would have been more reliable, if a stable boiling
regime was established during their experiments.
Previously, we utilized an oil heating method to heat the inside
of a tube shaped specimen and conducted electrochemical measurements on the outsider surface.25 The circulating oil heat eliminated
the signal noise caused by the electric heating method. A stable pool
boiling regime was reached and mass transfer rate was successfully
determined under the nucleate pool boiling condition. In this paper, a
novel flow boiling setup is designed and constructed. The flow boiling
system employed a similar oil heating method for follow up investigations based on our previous work.26 The experimental set up enable
us to study the effect of forced convection on mas transfer coefficient due to the micro-mixing and macro-mixing effects caused by
the generation and movement of bubbles, which has never studied
before. The mass transfer coefficients are determined experimentally
at elevated temperatures and pressures as well as various electrolyte
flow rates under a single-phase condition to verify the experimental
setup. The mass transfer coefficient of hydrogen peroxide on a boiling
surface electrode under a flow boiling condition are then obtained and
a new empirical correlation is proposed to represent the mass transfer
behavior under a flow boiling condition.
Experimental
Set up.To investigate the mass transfer behavior of dissolved
species in a flow boiling condition, a novel experimental setup is
designed and constructed. The schematic diagram of the flow loop
system and the electrochemical flow cell are shown in Figs. 1a and
1b, respectively. As shown in Fig. 1a, two stainless steel storage tanks
are used to store the chemistry-controlled electrolyte and to collect the
return electrolyte. A positive displacement pump (Neptune 525-S-N3)
with a maximum discharge pressure up to 6.2 MPa is used to circulate electrolyte solution and its flow rate is measured using a turbine
meter (Omega FTB504-CK). The inlet electrolyte temperature (T1 ) is
controlled using an electrical preheater. Then, the solution temperature is further heated to the boiling condition in the electrochemical
flow cell, where the electrochemical measurements are conducted using flowing hot oil. The heated oil flows between the heating bath
(Cole-Parmer SS 13L) and the inner tube of the flow cell while the
electrolyte flows through the annular space between the two tubes. Ktype thermocouples are used to measure the temperatures of the heated
oil before entering and after leaving the specimen area (T2 and T3 ).
The electrolyte is cooled to near room temperature using a shell and
tube heat exchanger (S.E.C. Heat exchangers Model C-8.2) after the
electrochemical flow cell. The pressure of the loop is controlled using
a backpressure regulator (Equilibra GSD4-SS316) and a check valve.
Fig. 1b shows the detailed schematic of the electrochemical flow
cell. The flow cell contains a double tube structure. The outer tube is
made of 316 Stainless Steel with an inner diameter of 15.75 mm and
an outer diameter of 19.05 mm. The inner tube is made of Alloy 800
heat exchanger tubing with an inner diameter of 7.80 mm, an outer
diameter 9.53 mm. The working electrode is 300 mm in length. The

Downloaded on 2016-05-16 to IP 129.128.135.134 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 163 (7) H618-H624 (2016)

(a)

Cool stream

Electrochemical
flow cell

Hot stream

H619

Union

T3

Oil flow

Oil heating
bath

Thermocouples

T1
Back pressure
regulator
Union

8. Heat

Solution
tank 2

T2

exchanger
Preheater

Chemistry
control

T4

Oxygen meter

Pump and
controller

Flow meter
Solution
tank 1
Check valve

Pressure relief valve

(b)
1. Working Electrode
Surface

3.1 Swagelock Adapter


1/4'' Female NPT to3/4'' Male NPT
SS-12-RB-4
2.1 Conax Fitting
1/4'' Male NPT
MHC2-062-B1-L

4. Union Cross 1''


SS-1610-4

5. Tube 1''

6. Swagelock Reducer
1'' Tube OD to 3/4'' Female NPT 7. Counter Electrode
SS-16-TA-7-12
Pt (ring)

9. Pressure Balanced
Ag/AgCl Reference Electrode
Working Fluid Flow

8. PTFE Shield

Oil Flow

Figure 1. (a) Schematic diagram of the flow loop and (b) the electrochemical flow cell.

measurements are conducted on the outer surface of the tube working


electrode where boiling occurs. This working electrode is located 78
cm away from of both ends of the electrochemical flow cell to avoid
any entrance length effect on the mass transfer behavior. A platinum
tube ring with a diameter of 15.5 mm and length 30 mm is used as a
counter electrode. A pressure balanced Ag/AgCl reference electrode
(Corr Instruments, LLC), can be operated up to 250 C is used to
provide constant potential during the measurement. All the potential
values and reported in this paper have been converted to the standard
hydrogen electrode (SHE).
Procedure.Before each test, the out surface of the working electrode was polished using #60 grit sand paper to obtain an average
roughness of 0.95 m so that the roughness was close to the conventional operating condition of heat exchanger tubes in the steam
generators. A 0.025 M Na2 SO4 solution was used as the supporting
electrolyte, and H2 O2 of 0.002 M dissolved in the solution was used

as the reaction species. Before each measurement, the electrolyte solution was deoxygenated by purging pure nitrogen gas for 2 hours.
The electrochemical measurements were conducted using a Gamry
Reference 600 electrochemical station. Potentiodynamic polarization
was applied by sweeping the potential from 50 mV above the Open
Circuit Potential (OCP) to the negative direction at rate of 2 mV/s until
the current density reached 1 mA/cm2 . The mass transfer coefficient,
k, and heat flux, q, were calculated using Eqs. 1 and 2, respectively.
k=

iL
n FCb

[1]

m C p (T2 T3 )
[2]
A
where iL is the measured limiting current density, n is the number
of electrons transferred per reaction, F is the Faraday constant, Cb is
q=

Downloaded on 2016-05-16 to IP 129.128.135.134 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

H620

Journal of The Electrochemical Society, 163 (7) H618-H624 (2016)

Table I. Diffusivities of hydrogen peroxide at various temperatures.


60
1.59 109

where i L is the total error in the measured limiting current density.


The value of i L was calculated from the sum of the systematic (accuracy) and random (precision) errors of the data. The accuracy error
comes from the maximum absolute error in the measuring device.
The maximum absolute error in the limiting current measurements
was 0.3%. The precision error was obtained directly from the standard deviation, , of the measured values. To determine the precision
error, Coleman and Steel27 stated that when the number of data points
for one time series was equal to or greater than 10, two times the standard deviation gave a good approximation for the 95% confidence
interval. Therefore, the total error in the measured limiting current
was calculated by
i L = (0.003i L + 2i L )

[4]

The total error in the bulk ion concentration, Cb , was assumed


to be only due to error in the preparation of electrolyte solution,
which was assumed to be 4%. Therefore, the error in the bulk ion
concentration was obtained by
Cb = (0.04Cb )

[5]

The mass transfer rate of H2 O2 was investigated under single-phase


flow and flow boiling conditions. Under the single-phase condition,
the experiments were conducted for the electrolyte temperature, flow
rate and operating pressure up to 120 C, 44.6 kg/m2 s and 652 kPa,
respectively. Under the flow boiling condition, the measurements were
conducted at flow rates up to 44.6 kg/m2 s, and oil temperatures up
to 160 C while the operating pressure was kept constant at 100 kPa.
The diffusivity values of hydrogen peroxide are also included in Table
I where the diffusivity at 20 C was obtained from literature29 and
the values for other temperatures were calculated using the StokesEinstein equation.30
Results and Discussion
Single-phase flow condition.Effect of temperature on mass
transfer coefficient.Figs. 2a and 2b show the polarization curves
and mass transfer coefficients for the hydrogen peroxide reaction as
a function of the solution temperature while the operating pressure
and electrolyte flow rate were kept constant at 445 kPa and 17 kg/m2
s, respectively. It can be observed from the Fig. 2a that the current
plateaus appear in the potential range between 1.15 to 1.35 V vs.
standard hydrogen electrode (SHE) for all electrolyte temperatures.
These plateaus indicate that the mass transfer controlled region can be
reached under the single-phase condition. It can also been seen that
the limiting current density (the current value of the plateau region)
increases with increasing temperature, which indicates that the mass
transfer controlled reaction rate increases. The mass transfer coefficients were then calculated using Eq. 1 and the measured limiting
current density obtained in the plateau region. Fig. 2b shows that the

100
2.89 109

120
3.54 109

mass transfer coefficient increases as the electrolyte temperature increases from 22 C to 120 C. Increasing the electrolyte temperature
increases the diffusivity of the dissolved hydrogen peroxide, therefore
the increase in the mass transfer coefficient is due to the increase in
diffusivity.29,31
Effect of the flow rate on the mass transfer coefficient.Fig. 3a
shows the polarization curves for the hydrogen peroxide reaction as
a function of electrolyte flow rates, while the operating pressure and
electrolyte temperature were kept constant at 445 kPa and 22 C, respectively. It can be observed from the Fig. 3a that there are clear
current plateaus in the potential range between 0.95 to 1.35 V
vs. SHE for all electrolyte flow rates. Fig. 3b shows the calculated
mass transfer coefficient as a function of the electrolyte flow rate for
various electrolyte temperatures while keeping the operating pressure
constant at 445 kPa. For all flow rates, the flow regime in the annular
area was in the laminar region (Re > 2100). It can be seen that the
mass transfer coefficient increases with the increasing electrolyte flow
rate for all electrolyte temperatures. A higher electrolyte flow rate induces a more intensive forced convection, thus, lead to a higher mass

1x10

-2

(a)

the bulk ion concentration, m is the electrolyte flow rate, Cp is the


heat capacity of the heating oil, T2 and T3 are the temperatures of the
heating oil before entering and after leaving and A is the surface area
of the working electrode.
The uncertainty in the experimentally determined mass transfer
coefficient, k was calculated using the method described by Colman
and Steel27 and Holman.28 The uncertainty in the experimentally determined mass transfer coefficient based on experimental errors in the
measured limiting current density, i L and the bulk ion concentration,
Cb , was calculated by

2 
2 0.5
k
k
k =
i
+
C
[3]
i L L
Cb b

80
2.21 109

Current Density, i (A/cm )

20
6.60 1010

Electrolyte Temperature ( C)
1
2

3
4

22
80

100
120

1x10

-3

2
1
2

Flow rate 17.0 kg/m s


Pressure 445 kPa
-4

1x10
-0.45

-0.70

-0.95

-1.20

-1.70 -1.85

-1.45

Potential, E (V vs. SHE)


Mass Transfer Coefficient, k (m/s)

Temperature ( C)
Diffusivities (m2 /s)

6.0x10

-5

(b)

4.0x10

-5

2.0x10

-5
2

Flow rate 17.0 kg/m s


Pressure 445 kPa

0.0
20

80

100

120

Electrolyte Temperature, T1 ( C)
Figure 2. (a) Cathodic polarization curves and (b) the mass transfer coefficients under the single phase flow condition at various electrolyte temperatures.

Downloaded on 2016-05-16 to IP 129.128.135.134 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 163 (7) H618-H624 (2016)


-3

1x10

-3

8.0x10

(a)
o
Electrolyte Temperature 22 C
Pressure 445 kPa
5
4

5x10

-4

2
1

Electrolyte Flow Rate (kg/m s)

-4

1x10
-0.45

-0.70

-0.95

1
2
3
4
5

8.9
12.2
17.0
29.2
44.6

-1.20

-1.45

Mass Transfer Coefficient, k (m/s)

Current Density, i (A/cm )

2x10

-5

(b)
o
Electrolyte Temperature ( C)
6.0x10

Flow Rate 17.0 kg/m s


4.0x10

-5

2.0x10

-5

0.0

-1.70 -1.85

300

Mass Transfer Coefficient, k (m/s)

Electrolyte Temperature ( C)

-5

22
80
-5

4.0x10

-5

2.0x10

-5

900

Figure 4. The mass transfer coefficients as a function of operating pressure


and for various electrolyte temperatures.

-4

6.0x10

600

Pressure (kPa)

(b)
8.0x10

100
120

22
80

-5

Potential, E (V vs. SHE)


1.0x10

H621

ter, which leads to increase in the diffusion coefficient.35 As a result,


the mass transfer coefficient increases with increasing operating pressure for all electrolyte temperatures. The findings of the present study
are in agreement with the model and experimental results given by
Al-Dahhan et al.36,37

100
120

Pressure 445 kPa

0.0

10

20

30

40

50

Electrolyte Flow Rate (kg/m s)

Correlation for single-phase flow condition.Mass transfer under


a single-phase condition has been well studied in the open literature.
Lin et al.38 examined the mass transfer rate for the cathodic reaction
of ferricynaide ions from a flowing electrolyte to the inside surface
in an annular shaped tubing. Their results agreed well with Leveques
equation in the laminar flow region and with Chilton-Colburns equation in the turbulent flow region.32 Carbin and Gabe39 further studied
the mass transfer behavior though the annular geometry based on Lin
et al.s study.38 Carbin and Gabe39 proposed the following correlation
to predict the mass transfer coefficient under the single-phase condition, ksp , in the range of 100 < Re < 50000 and 100 < Sc < 2000:


Figure 3. (a) Cathodic polarization curves and (b) the mass transfer coefficients as a function of electrolyte flow rate for various electrolyte temperatures.

transfer coefficient. These trends agree well with the classic Leveque
equation.32
Effect of the operating pressure on the mass transfer rate.The
polarization curves were also obtained for various operating pressures
while the electrolyte temperature and the flow rate were kept constant.
In these tests, the operating pressure was varied from 100 to 652 kPa
and for each operating pressure, the electrolyte temperature was varied from 22 to 120 C while the electrolyte flow rate was kept constant
at 17.0 kg/m2 s. For all operating pressures and the electrolyte temperatures, the mass transfer controlled reaction region was reached.
Fig. 4 shows the mass transfer coefficients as a function of operating pressure for various electrolyte temperatures. The data points at
the operating pressures of 100 and 238 kPa and electrolyte temperatures of 100 and 120 C were not included because the electrolyte
temperature exceeded the saturation point and the flow was no longer
single-phase. It can be seen from Fig. 4 that under single-phase flow
condition, the mass transfer coefficient is slightly increased when the
operating pressure increased. The increase of pressure enhanced the
inner energy of the electrolyte solution, thus increasing the movement of dissolved hydrogen peroxide.33 The Gibbs free energy, which
atoms need to perform a diffusion jump is dependent on pressure.34
An increase in pressure decreases the activation volume of water as
well as breaks down the tetrahedrally coordinated structure of wa-

Sh = 3.93 Re0.32 Sc0.33

dc
Le

0.35
[6]

Therefore, the single-phase mass transfer coefficient was expressed


as:
ksp = 3.93

D
Re0.32 Sc0.33
dc

dc
Le

0.35
[7]

where D is the diffusivity, dc is the hydraulic diameter and Le is


the length of the electrode. They stated that their correlation agreed
well with others work.12,38 Therefore, Eq. 7 was used to predict the
mass transfer coefficients under the single-phase condition to verify
the experimental data of the present study. Fig. 5 shows the comparison between the experimental data and the predicted mass transfer
coefficients obtained using the Eq. 7. It can be observed that 95% of
the experimentally determined mass transfer coefficients agree with
the correlation developed by Carbin and Gabe39 within 15% error
(dashed lines) under the single-phase flow condition. This agreement
confirms that the flow loop constructed for the present produces reasonable mass transfer data and the electrochemical method is the
appropriate measuring technique to determine the mass transfer coefficient under a single-phase flow condition.
Mass transfer measurement under the flow boiling condition.
Initially, the potential static test was conducted by recording the limiting current density and the oil temperatures as a function of time
to determine if the fully developed flow boiling regime occurred in
the annular space between the two tubes. Fig. 6 shows the limiting

Downloaded on 2016-05-16 to IP 129.128.135.134 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 163 (7) H618-H624 (2016)


-4

0.04

1.0x10

Current Density, i (A/cm )

Predicted Mass Transfer Coefficient, k (m/s)

H622

+15%

-5

8.0x10

-5

6.0x10

-15%

-5

4.0x10

Equation (7)
15% error range
Experimental Data

-5

2.0x10

0.0
0.0

1x10-2

(b)
Flow boiling condition
2
Heat flux (kW/m )
1
2
3

223.06
250.00
290.42

1
-3

1x 10

Single phase condition

-4

-5

-5

2.0x10

4.0x10

-5

6.0x10

-5

8.0x10

-4

1.0x10

Experimental Mass Transfer Coefficient, k (m/s)

1x10
-0.45

-0.70

-0.95

-1.20

-1.45

-1.70 -1.85

Potential, E (V vs. SHE)

Figure 5. The comparison between the experimentally determined and predicted mass transfer coefficients under single phase flow condition with the
two dashed lines the error range.

Figure 7. Cathodic polarization curves for single-phase and flow boiling


conditions.

current density measured at the constant potential of 1.3 vs. SHE


as a function of time when the mineral oil temperature was increased
from room temperature up to 150 C. The sudden increase in the limiting current density at 700 s is attributed to the phase changing in
the electrolyte solution when the oil temperature was around 110 C.
The reason for this sudden increase is attributed to the initiation and
expansion of bubbles during boiling enhancing the forced convection
on the specimen and bubble surfaces. Between 700 s and 1200 s,
the limiting current density continues to increase with the increasing
oil temperature. The reason for this increase may be due to the flow
regime changes from the subcooled flow boiling to the fully developed
flow boiling condition. After 1200 s, the limiting current density and
the oil temperatures remain relatively constant. This indicates that the
fully developed flow boiling regime is reached and the mass transfer
controlled reaction has a stable reaction rate region.
Fig. 7 shows the cathodic polarization curves for the single-phase
and flow boiling conditions. Clear current plateaus can be observed in
the potential range of 1.20 V to 1.40 V vs. SHE for all heat fluxes.
It can be seen that the current densities and consequently the mass
transfer reaction rates under the flow boiling condition are higher than
that of the single-phase condition. There are two reasons for this; the
first reason is the surface electrolyte temperature under the flow boiling condition is higher than that under the single-phase condition and
the other is the effect of bubbles during boiling. The micro-mixing

and macro-mixing effects caused by the generation and movement


of bubbles enhanced the forced convection. The bubbles also break
the diffusion layer and reduced the concentration gradient.40,41 The
combined effect is an increase in the mass transfer rate. Fig. 7 also
shows that the limiting current density values increase with increasing heat flux under the flow boiling condition. The increase in heat
flux increases the vapor generation rate, which means more bubbles
are generated on the specimen surface during boiling. The movement
generated by bubbles expanding and detaching can further break the
diffusion layer. Consequently, the forced convection increases and
brings more reactive compounds to the interface where the concentration of reactive compounds was zero. This effect remarkably increases
the mass transfer rate from the bulk solution to the specimen interface.
Empirical correlation for mass transfer behavior under the flow
boiling condition.The mass transfer correlations were largely obtained from heat transfer due to the boundary layer similarity.42 One of
the classic correlation for the flow boiling heat transfer coefficient was
developed by Chen.43 The heat transfer coefficient is the summation
of the heat transfer coefficients contributed by both the pool boiling
effect and the flow rate effect. This correlation has been further refined
by Gungor and Winterton44 for applications involving both vertical
and horizontal flows under saturated and subcooled boiling conditions
shown as follows:
h f b = E 1 h pool + E 2 h sp

0.004
2

(a)
T3

0.003

Limiting current
density

Temperatures, T ( C)

T2
120

0.002
80
0.001
Developing

40

Limiting Current Density, iL (A/cm )

160

300

600

900

Time, t (s)
Figure 6. Potentiostatic test as a function of time.

Flow
Boiling

1200

0.000
1500

[8]

where h is the heat transfer coefficient, the subscripts pool and sp


refer to pool boiling and single-phase forced convection, respectively.
The variable E1 is the suppression factor and E2 is the enhancement
factor. The following Dittus-Bolter45 and Cooper46 correlations can
be used to determine the heat transfer coefficients contributed by the
pool boiling and the single-phase forced convective terms:

0.55 0.5 0.67
M
q
[9]
h pool = 55Pr 0.12 log10 Pr

h sp = 0.023 Re0.8 Pr 0.4


d

[10]

where Pr is the Prandtl number, M is the molecular weight, q is the


heat flux, is the thermal conductivity and d is the tubing diameter. However, the existing heat transfer correlations mentioned above
under the flow boiling condition cannot be used for the prediction of
the mass transfer coefficient due to the boundary layer discontinuity.
According to open literature, there is no existing correlation available for predicting the mass transfer coefficient under the flow boiling
condition. However, there are few mass transfer correlations for the

Downloaded on 2016-05-16 to IP 129.128.135.134 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Experimental Mass Transfer Coefficient,k, (m/s)

Journal of The Electrochemical Society, 163 (7) H618-H624 (2016)

H623

1000 and a heat flux from 220 kW/m2 to 280 kW/m2 . Since there is
no available data under flow boiling condition in the open literature,
we were not able to compare the results from this paper with others
work.

-4

1.0x10

-5

9.0x10

+10%
-5

8.0x10

Conclusions

-5

7.0x10

-10%
-5

6.0x10

Equation (13)
-5

10% error range

5.0x10

Experimental data
-5

4.0x10
-5
4.0x10

-5

5.0x10

-5

6.0x10

-5

7.0x10

-5

8.0x10

-5

9.0x10

-4

1.0x10

The mass transfer coefficients of the hydrogen peroxide were successfully obtained experimentally using an electrochemical method
at elevated temperatures and pressures for both single-phase flow and
flow boiling conditions. Under the flow boiling condition, the electrochemical method was successfully employed to determine the mass
transfer coefficient for the first time. A new mass transfer correlation
was proposed and found to be valid to represent the experimental data
within 10% of error for the flow boiling condition. This correlation
filled the knowledge gap for the mass transfer under a flow boiling
condition and provides a reference for the future studies.

Predicted Mass Transfer Coefficient, k, (m/s)


Figure 8. The comparison between the experimentally determined and predicted mass transfer coefficient under flow boiling condition with the two
dashed lines the error range.

single-phase annular flow39 and pool boiling conditions.25 According to Carbin and Gabes39 study, the mass transfer coefficient for the
single-phase annular flow condition, ksp has been shown in Eq. 7. Shen
et al.25 developed the following equation for the mass transfer coefficient under the pool boiling condition, kpool , including micro-mixing,
macro-mixing and non-isothermal convection effects.


2.09 VG db D 0.5 0.01 D Ra y 0.5 0.012 D (Sc Gr )0.33
k pool =
+
+
dc
A
dc
dc
[11]
where VG is the gas evolution rate, db is the break off diameter of
the gas bubbles, Ray is the Rayleigh number and Gr is the Grashof
number.47,48
Using the same idea as the heat transfer analogy, the following
correlation is proposed to predict the mass transfer coefficient for the
flow boiling condition, kfb . The proposed correlation is the summation
of the mass transfer coefficient due the flow rate and pool boiling
effects:
k f b = F1 k pool + F2 ksp

[12]

where kpool is the mass transfer coefficient due to pool boiling and ksp
is the mass transfer coefficient under the single-phase flow condition.
The variable F1 is the suppression and F2 is the enhancement factors.
The suppression, F1 and enhancement, F2 factors in Eq. 12 were
obtained using a linear regression algorithm in conjunction with Eqs.
7 and 11 and with 2/3 of the flow boiling mass transfer experimental
data. The following correlation was obtained to describe the flow
boiling mass transfer behavior:
k f b = 0.699 k pool + 1.186 ksp

[13]

In Eq. 13 the constant F1 is smaller than unity, which indicates


that the bubble-induced enhancement of mass transfer in the flow
boiling environment is not as significant as the effect under the pool
boiling condition. But the constant for the mass transfer due to forced
convection mass transfer, F2 is higher than unity, which indicates
that the single-phase forced convection mass transfer under the flow
boiling condition, has significant effect on the two-phase mass transfer
rate, ktp .
Fig. 8 shows the comparison between the experimental mass transfer coefficients and predicted line calculated using proposed Eq. 13
for the flow boiling condition. It can also be seen that Eq. 13 represents all experimentally determined mass transfer coefficients within
10% (dashed lines) under flow boiling in annuli flow. The proposed
equation is valid within the Reynolds number, Re range of 200 and

References
1. S. Krishnamurthy and Y. Peles, International Journal of Heat and Mass Transfer, 51,
1349 (2008).
2. S. L. Qi, P. Zhang, R. Z. Wang, and L. X. Xu, International Journal of Heat and
Mass Transfer, 50, 4999 (2007).
3. N. Basu, G. R. Warrier, and V. K. Dhir, Journal of Heat Transfer, 127, 141 (2005).
4. L. Zhang, Z.-D. Li, K. Li, H.-X. Li, and J.-F. Zhao, The Journal of Computational
Multiphase Flows, 6, 313 (2014).
5. C. S. Brooks, B. Ozar, T. Hibiki, and M. Ishii, Nuclear Engineering and Design, 268,
152 (2014).
6. K. Khumsa-Ang and D. Lister, Heat Transfer Engineering, 34, 702 (2013).
7. A. Tvarusko, Journal of The Electrochemical Society, 120, 87 (1973).
8. M. W. Losey, M. A. Schmidt, and K. F. Jensen, Industrial & Engineering Chemistry
Research, 40, 2555 (2001).
9. D. R. Diercks, W. J. Shack, and J. Muscara, Nuclear engineering and design, 194,
19 (1999).
10. M. Al-Dahhan, W. HighFill, and B. T. Ong, Industrial & Engineering Chemistry
Research, 39, 3102 (2000).
11. J. R. Selman and C. W. Tobias, Advances in Chemical Engineering, 10, 211 (1978).
12. W. L. Friend and A. Metzner, AIChE Journal, 4, 393 (1958).
13. T. K. Ross and A. A. Wragg, Electrochimica Acta, 10, 1093 (1965).
14. A. A. Wragg, Electrochimica Acta, 13, 2159 (1968).
15. A. A. Wragg and A. K. Nasiruddin, Electrochimica Acta, 18, 619 (1973).
16. S. Gedupudi, Y. Q. Zu, T. G. Karayiannis, D. B. R. Kenning, and Y. Y. Yan, International Journal of Thermal Sciences, 50, 250 (2011).
17. M. A. Hakeem and M. Kamil, Chemical Engineering Research and Design, 85, 1670
(2007).
18. O. Zeitoun and M. Shoukri, International Journal of Heat and Mass Transfer, 40,
869 (1997).
19. H. Vogt, O. Aras, and R. J. Balzer, International Journal of Heat and Mass Transfer,
47, 787 (2004).
20. H. Vogt, International Journal of Heat and Mass Transfer, 59, 191 (2013).
21. A. Beckmann, B. A. Coles, G. Richard, P. Grandler, F. Marken, and A. Neudeck, The
Journal of Physical Chemistry B, 104, 764 (2000).
22. A. Beckmann, A. Schneider, and P. Grundler, Electrochemistry Communications, 1,
46 (1999).
23. F. Marken, S. L. Matthews, R. G. Compton, and B. A. Coles, Electroanalysis, 12,
267 (2000).
24. H. Sarac, A. Wragg, and M. Patrick, International Communications in Heat and Mass
Transfer, 30, 909 (2003).
25. C. Shen, A. Afacan, J.-L. Luo, A. Siddiqui, and S. J. Klimas, International Journal
of Heat and Mass Transfer, 78, 289 (2014).
26. C. Shen, A. Afacan, J.-L. Luo, A. Siddiqui, and S. J. Klimas, Corrosion, 71, 1003
(2015).
27. H. W. Coleman and W. G. Steele, Experimentation and uncertainty analysis for
engineers, John Wiley & Sons Incorporated (1999).
28. J. P. Holman, Experimental Methods for Engineers, McGraw Hill (2001).
29. S. B. Hall, E. A. Khudaish, and A. L. Hart, Electrochimica Acta, 43, 579 (1998).
30. J. T. Edward, Journal of Chemical Education, 47, 261 (1970).
31. S. B. Hall, E. A. Khudaish, and A. L. Hart, Electrochimica Acta, 44, 2455 (1999).
32. F. P. Berger and K. F. F. L. Hau, International Journal of Heat and Mass Transfer,
20, 1185 (1977).
33. G. Schock and A. Miquel, Desalination, 64, 339 (1987).
34. H. Mehrer, in Defect and Diffusion Forum, p. 57 (1996).
35. R. Cuddeback, R. Koeller, and H. Drickamer, The Journal of Chemical Physics, 21,
589 (1953).
36. W. Highfill and M. Al-Dahhan, Chemical Engineering Research and Design, 79, 631
(2001).
37. M. H. Al-Dahhan, F. Larachi, M. P. Dudukovic, and A. Laurent, Industrial & engineering chemistry research, 36, 3292 (1997).

Downloaded on 2016-05-16 to IP 129.128.135.134 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

H624

Journal of The Electrochemical Society, 163 (7) H618-H624 (2016)

38. C. S. Lin, E. B. Denton, H. S. Gaskill, and G. L. Putnam, Industrial & Engineering


Chemistry, 43, 2136 (1951).
39. D. C. Carbin and D. R. Gabe, Electrochimica Acta, 19, 653 (1974).
40. R. C. Engstrom, M. Weber, D. J. Wunder, R. Burgess, and S. Winquist, Analytical
Chemistry, 58, 844 (1986).
41. L. J. J. Janssen and J. G. Hoogland, Electrochimica Acta, 15, 1013
(1970).
42. R. Ellahi, M. M. Bhatti, and K. Vafai, International Journal of Heat and Mass
Transfer, 71, 706 (2014).

43. J. C. Chen, Industrial & Engineering Chemistry Process Design and Development,
5, 322 (1966).
44. K. E. Gungor and R. H. S. Winterton, International Journal of Heat and Mass Transfer, 29, 351 (1986).
45. F. W. Dittus and L. M. K. Boelter, International Communications in Heat and Mass
Transfer, 12, 3 (1985).
46. M. Cooper, in Inst. Chem. Eng. Symp. Ser, p. 785 (1984).
47. K. Stephan and H. Vogt, Electrochimica Acta, 24, 11 (1979).
48. G. Son and V. Dhir, Journal of Heat Transfer, 120, 183 (1998).

Downloaded on 2016-05-16 to IP 129.128.135.134 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Potrebbero piacerti anche