Sunteți pe pagina 1din 5

Powder Technology 190 (2009) 385–389

Contents lists available at ScienceDirect

Powder Technology
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / p o w t e c

Mass transfer in the core-annular and fast fluidization flow regimes of a CFB
Ronald W. Breault ⁎, Christopher P. Guenther
US Department of Energy, National Energy Technology Laboratory, Morgantown, WV 26507, USA

a r t i c l e i n f o a b s t r a c t

Article history: In gas–solid reactors, particularly circulating fluidized beds (CFB) it is becoming increasingly more important to
Received 21 May 2008 be able to predict the conversion and yield of reactant species given the ever rising cost of the reactants and the
Received in revised form 11 August 2008 ever decreasing acceptable level of effluent contaminants. As such, the development and use of predictive
Accepted 17 August 2008
models for the reactors is necessary for most processes today. These models all take into account, in some
Available online 27 August 2008
manner, the interphase mass transfer. The model developer, unless equipped with specific experimentally
Keywords:
based empirical correlations for the reactor system under consideration, is required to go to the open literature
Mass transfer to obtain correlations for the mass transfer coefficient between the solid and gas phases. This is a difficult task at
Multiphase flow present, since these literature values differ by up to 7 orders of magnitude. The wide variation in the prediction
Circulating fluidized bed of mass transfer coefficients in the existing literature is credited to flow regime differences that can be
Clusters identified in the cited literature upon careful inspection.
Fluctuations A new theory is developed herein that takes into account the local hydrodynamics. The resulting model is
compared with data generated in the NETL cold flow test facility and with values from the literature. The new
theory and the experimental data agree quite well, providing a fundamentally based mass transfer model for
predictive reactor simulation codes.
Published by Elsevier B.V.

1. Introduction lating fluidized beds which is available in greater detail in a review


conducted by Breault [6]. The second part is a discussion of mass
A major need of the system designers for the simulation of complex transfer theory in liquid films that will be used to develop the
gas–solid chemically reacting flows such as those occurring in analogous theory for the solids in a CFB.
circulating fluidized bed systems is accurate, hydrodynamic based
transport relationships for mass transfer, dispersion and heat transfer 2.1. Mass transfer in circulating fluidized beds
by analogy. This was demonstrated in work by Breault and Guenther
[1,2]. That undertaking showed that the model predictions are A review of the literature revealed six experimental investigations
clearly sensitive to both the value of the Sherwood number and the on mass transfer in circulating fluidized beds (Li [7], Subbarao [8],
value of the gas/solids Peclet numbers. Breault et al. [3] presented an Bolland [9], Kalil [10], Resnick [11], and Venderbosch [12]). A de-
experimental method to develop local dispersion coefficients from scription of the operating conditions and experimental facilities for
time dependent velocity data at a point in the riser. This methodology the review of the mass transfer data is presented in Table 1. These
was utilized by Jildarak et al. [4,5] to show that the technique could studies were generally conducted in small diameter, bench-top
be used with CFD generated data to obtain the local dispersion facilities; the exception being the work of Bolland [9]. Also, the bulk
coefficients. Therefore, a method exits to use information that is al- of the work was conducted at ambient conditions or near ambient
ready being predicted within the simulation to calculate the local conditions with the only exception being the work of Vanderbosch
dispersion coefficients. Relationships are needed to predict the [12], which was conducted at 750 K.
interphase mass transfer and gas–solid heat transfer in these chem- The correlations developed from the six experimental investi-
ically reacting circulating fluidized bed systems. gations noted above are presented graphically in Fig. 1 with the
corresponding mathematical expression given in Table 2. The cor-
2. Literature review relations of two additional researchers (Gunn [13] and Zennerhoven
[14]) are also presented in this figure. These latter two are based on
The following literature review is summarized into two parts. reviews of other's data. There is little agreement between the re-
The first part is a short summary of mass transfer literature on circu- ported results with values of the Sherwood number ranging from a
low of 10− 5 reported by Bolland [9] to a high of 200 as reported by Kalil
[10].
⁎ Corresponding author. Looking at the specific research that was conducted, it is possible to
E-mail address: Ronald.Breault@NETL.DOE.gov (R.W. Breault). reduce this span by a couple orders of magnitude. Bolland conducted

0032-5910/$ – see front matter. Published by Elsevier B.V. ‫ﻣﺤﺪوده‬


doi:10.1016/j.powtec.2008.08.021
386 R.W. Breault, C.P. Guenther / Powder Technology 190 (2009) 385–389

Table 1
Conditions and dimensions for mass transfer experimental investigations

Author d (m) h (m) T (K) P (bar) dp (um) ρs (kg/m3) Gs (kg/m2 s) Ug (m/s)


Li 0.072 3 Ambient Ambient 300 1200 6 and 12 1.86 and 2.12
Subbarao 0.025 1.05 Ambient Ambient 196 and 390 2500 ∼15 to 60 4, 5 and 6
Bolland 0.411 8.5 333 Ambient 134 2500 31 to 53 5.6 to 7.2
Kalil 0.09525 0.3 Ambient Ambient 685 to 2000 800 to 10,000 N/A N/A
Resnick .022 to .044 0 Ambient Ambient 3 to 35 mesh 0 N/A N/A
Venderbosch 0.015 1.17 750 Ambient 65 1375 5 to 40 2.5 to 4.5

Bolland Correlation is highly suspected Kalil correlation is suited for bubbling fluidized bed not CFB
his research utilizing an ozone decomposition reaction. Ozone, upon have been put forward to explain the phenomena on a fundamental
first thought seems to be a good tracer to use, after all it decomposes basis. These are – in increasing complexity – penetration theory
to oxygen. Ozone decomposition is also highly dependent upon the (Higbie [15]), penetration theory with random surface renewal
moisture concentration in that gas and as no moisture data was taken (Dancwerts [16]), film-penetration theory (Toor and Machello [17])
and no effort made to control the moisture content during Bolland's and transfer to falling turbulent wavy films (Banerjee et al. [18]). In all
test, his reported correlation for the Sherwood number is highly of these theories, the mass transfer coefficient is proportional to the
suspected. Eliminating this correlation from the analysis tightens the square root of the diffusivity and inversely proportional to the square
data range substantially, now being 0.01 to 200. The work of Kalil [10] root of a time parameter which represents the contact time.
was conducted in a bubbling fluidized bed. He reports that the bed
voidage is an important factor in the correlation of the Sherwood sffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffi
number. However, the application of this correlation to the riser of a 4D pffiffiffiffiffiffi 4Du
k≈ ≈ sD ≈ ð2Þ
CFB where voidages are on the order of 0.8 and higher likely πte πδ
introduces significant error in the estimate of the Sherwood number
as the voidage in Kalil's work was on the order of 0.6. Thus, it can be
where
expected for the Sherwood number to fall between 0.01 and 10. This is
still a significant difference, but no other factors could be identified
1 u
which suggest that any of the other correlations should be discarded. ¼ is ð3Þ
te δ
Looking at the values of the Sherwood number in Fig. 1, leads to
questioning the validity of using a modified single sphere expression
of the form where te is the exposure time, u is the eddy velocity, δ is the eddy size
and s is the fractional rate of surface renewal.
Sh or Nu ¼ 2 þ aReb Scc ð1Þ
3. Development of mass transfer coefficient for a cluster in a CFB
In addition to the literature data presented and discussed above,
limited experimentation for both heat and mass transfer in the core- In analyzing the phenomena of mass transfer in a riser operating
annular flow regime has provided heat and mass transfer results under core-annular conditions or fast fluidization, we will first in-
that are less than single phase flow past a single sphere at the same troduce the concept that clusters are small packages of particles
gas velocity. This peculiar behavior combined with the need to model that move about in a random manner not significantly different than
and simulate the performance of these two reactors has induced the the movement of eddies in a turbulent single phase system, i.e. the
authors to reexamine the transport phenomena from a more analogy to Banerjee et al.'s work on turbulent wavy films (Banerjee
fundamental approach. et al. [18]). Visualize the system to consist of clusters moving along
a plane as shown in Fig. 2. In the model, clusters are made up of
2.2. Mass transfer in thin films particles moving together as a relatively dense suspension, possibly
approaching a packed bed density which have very little internal
For mass transfer into a liquid film, which moves down along a mass transport. That is to say that the clusters move with the fluc-
wall, from a gas containing a soluble species, a number of theories tuating velocity, uc, collide with one another to generate new sur-
face area at a time scale shorter than that for diffusion through the
quasi-static interial region of the clusters. These clusters have an
average cluster size of δ, the measurement of the length width and
height.
Given this model, the governing differential equation for mass
transfer from the bulk to the cluster can easilily be manipulated into a

Table 2
Mass transfer coefficient correlations

Author (PI) Model and/or data range


Resl 0:5 0:3
Li Sh ¼ 2e þ 0:69 Sc

e 
U ρ 1:43
Subbarao Sh ¼ 8:314⁎10−5 G0 s p
Bolland 10− 5 b Sh b 10− 3
Zevenhoven Sh ¼ 2e þ 0:89Re0:5 Sc0:33
 0:56p 0:33
Kalil Sh ¼ 1:77 Ree1−e Sc
Resnick Graphical J factor
Venderbosch Sh decreases with solids fraction and increase with U0
   
Gunn Sh ¼ 7−10e þ 5e2 1 þ 0:7Re0:2 p Sc
0:33
þ 1:33−2:4e þ 1:2e2 Rep0:7 Sc0:33
Fig. 1. Mass transfer coefficients [6].
R.W. Breault, C.P. Guenther / Powder Technology 190 (2009) 385–389 387

Fig. 2. Illustration of cluster model.

generic form and the solution obtained in any number of texts, for
example see Bird et al. [19], to be

x
C ¼ 1− erf pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð4Þ
4Dδ=uc

From this the mass transfer coefficient can be found to be


rffiffiffiffiffiffiffiffiffiffiffi
4Duc
k¼ ð5Þ
πδ
which is identical with the relationship developed by Banerjee et al.
Fig. 3. CFB Schematic.
[18] except that the velocity term is the cluster velocity and the length
term, δ, now represents the cluster size. As mentioned above, the diameter and 0.27-m above the gas distributor. Solids exit the riser
surface of the cluster is renewed through the interaction of adjacent through a 0.20-m port at 90° about 1.2-m below the top of the riser at a
clusters. Recalling that the cluster velocity term, uc, is the fluctuating point 15.45-m above the solids entry location (centerline to centerline).
velocity and the square of this velocity is the turbulent kinetic energy. Riser velocities were corrected for temperature and pressure as mea-
Therefore, the mass transfer coefficient can be expressed as sured at the base of the riser. The air's relative humidity was maintained
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi at 20% to minimize effects of static charge building up on the solids.
u qffiffiffiffiffiffiffiffiffiffiffi
u 2 Experiments were carried out by inserting a naphthalene sphere
t4D ðuc Þ
k¼ ð6Þ (moth ball) into the riser for a fixed time increment by holding the
πδ
mothball in the probe shown in Fig. 4. With the mothball securely
Techniques for determining cluster size and velocity have been clamped in the probe, it could be inserted and removed at will though
recently been reported by Breault et al. [20] and Guenther and Breault a series of ball valves. The weight loss divided by the exposed time
[21]. Breault et al. [20] using a statistical approach for determining the provided the mass transfer rate data needed to calculate the mass
existence of a cluster from LDV data found that clusters consist of a transfer coefficient.
minimum of six particles having a maximum particle separation equal
to the particle diameter. Therefore, the minimum cluster characteristic 5. Results
length can be taken to be 11 particle diameters. Guenther and Breault
[21] using wavelet analysis techniques of fiber optic time series data The data are plotted in Figs. 5 and 6 against the square root of the
found that cluster frequency and size were functions of the solids and Reynolds number. The data are then compared to the Gunn correlation
gas flow rates. Summarizing those results at constant gas velocity, the
cluster frequency increases 25% when the solids flow rate increases by
a factor of 5 while the cluster size increases by a factor of 3 for the
same increase in the solids flow rate. Also, at constant solids flow rate,
the cluster frequency is relatively independent of the gas velocity for
nearly doubling the gas velocity while the cluster size decreases by
about 30% for this same change in the gas velocity.

4. Experimental system

The test unit configuration is shown in Fig. 3 and also described


elsewhere [3,20]. The solids enter the riser from a side port 0.23-m in Fig. 4. Mass transfer probe.
388 R.W. Breault, C.P. Guenther / Powder Technology 190 (2009) 385–389

Table 3
Comparison of experimental and theoretical mass transfer coefficients

Mass transfer coefficient, m/s


Experimental Theory (Eq. (6))
Value Error Value Error
0.010 ±.002 0.012 +0.0004, −0.001
0.015 ±.003 0.014 +0.0008, −0.001

clusters will decrease and thereby result in the inverse relationship.


With this assumption, the result being that fluctuating solids energy
decreased by 20% as the solids flow rate increase by a factor of 5,
the Subbarao data point for a gas velocity of 5 m/s and a solids flux of
30 kg/m2 s was fitted to give the fluctuating solids kinetic energy. For
this value, the fluctuating solids energy was calculated for the solids
Fig. 5. Effect of solids flow on Sherwood numbers. flux rage of Subbarao's work. The results are plotted in Fig. 7. Also in
Fig. 7, Subbarao data at a gas velocity of 4 m/s is also presented. The
as presented in Table 2 as this correlation is the correlation presently corresponding model results are obtained by noting that the cluster size
used in the MFIX CFD code [22]. The gross assessment of the is a function of the gas velocity as reported by Guenther and Breault [21]
correlation to represent the data is that the correlation is reasonable, and discussed above and noting that for the change in gas velocity from
providing Sherwood numbers that are within acceptable error limits 5 m/s to 4 m/s the solids size would change by a factor of 1.87.
for mass transfer. The correlation appears to do better for higher To further test the theory, additional data by Vanderbosch [12]
Reynolds numbers than for low. In the CFB and transport reactor, the were also compared. Since these data were developed under different
Reynolds number is more likely to be less than 100. While the Gunn hydrodynamic conditions than the Subbarao data, the fluctuating
correlation provides a reasonable estimate of the Sherwood number, solids kinetic energy was fitted for the lowest solids flux value. The
additional data for Reynolds number less than 100 could provide some trends in the cluster size and the fluctuation solids kinetic energy with
useful information that would result in a better correlation. solids flow rate were maintained the same as before. The results of the
Table 3 presents a comparison of the experimental mass transfer model prediction are shown in Fig. 8 along with the earlier results for
coefficients with the theoretical values as calculated using Eq. (6). The the Subbarao data.
value for the average cluster size, δ, and the fluctuating cluster kinetic
energy, u2c , are obtained from fiber optic data analyzed and reported on
by Guenther and Breault [21]. The experimental values plotted are the
average values for the data at each condition. The error represents the
range in the data at each condition. The error on the theoretical values
are calculated using the average cluster size moving up and the average
cluster size moving down to give a wide range for these values.
To provide a further test and validation of the developed theory, the
literature was reviewed with the purpose of finding mass transfer data
with sufficient discussions regarding the hydrodynamics such that the
fluctuating cluster kinetic energy and the cluster size could be obtained
or estimated. The work by Sabbarao [8] provided such information. For
this analysis, the cluster size was taken as 11 particle sizes for the low
solids flow rate. Based upon the results of Guenther and Breault [21],
this size was allowed to increase linearly with solids flow rate such that
the cluster size increased by a factor of 3 for a solids flow rate increase
by a factor of 5. The cluster frequency was then assumed to inversely
correlate with the fluctuating solids kinetic energy. This assumption
Fig. 7. A comparison of Subbarao data with theoretical model predictions.
makes sense as one can visualize that as more and more clusters appear
with time, the flow is getting denser and the random velocity of the

Fig. 6. Comparison of data with Gunn correlation. Fig. 8. Comparison of Vanderbosch and Subbarao data with the theoretical predictions.
R.W. Breault, C.P. Guenther / Powder Technology 190 (2009) 385–389 389

6. Conclusions References

A theoretical based model was developed for predicting mass [1] R.W. Breault, C. Guenther, Sensitivity of gas–solids dispersion and mass transfer
transfer coefficients using an analogy that clusters are like turbulent coefficient in an Eulerian–Eulerian CFD modeling. Spring AIChE Meeting, Atlanta,
April 2005.
eddies. With this basic assumption, the governing differential equation [2] R.W. Breault, C. Guenther, Two and three dimensional simulations investigating
was set and solved to give a relationship for the mass transfer coefficient. dispersion and mass transfer coefficients in an Eulerian–Eulerian CFD model,
This relation ship was tested with experimental data obtained at NETL Fourth International Conference on Computational Fluid Dynamics in the Oil and
Gas, Metallurgical & Process Industries, 2005, Trondheim, Norway.
and with data in the literature and shown to give good predictive values [3] R.W. Breault, C. Guenther, L.J. Shadle, Velocity fluctuation interpretation in the
for widely different systems. near wall region of a dense riser, Powder Technology vol. 182 (Issue 2) (February 22
2008) 137–145.
[4] V. Jildarak, D. Gidaspow, R.W. Breault, Computation of gas and solid dispersion
Nomenclature coefficients in turbulent risers and bubbling beds, Chemical Engineering Science
C Concentration, mol/m3 vol. 62 (Issue 13) (July 2007) 3397–3409.
d diameter, m [5] V. Jildarak, D. Gidaspow, R.W. Breault, C. Guenther, L.J. Shadle, S. Shi, Computation
of turbulence and dispersion of cork in the NETL riser, Chemical Engineering
G Gas flux, kg/m2 s
Science (January 24 2008) Available online.
Gs Solids flux, kg/m2 s [6] R.W. Breault, A review of gas–solid dispersion and mass transfer coefficient
k rate constant or coefficient, m/s correlations in circulating fluidized beds, Powder Technology vol. 163 (Issues 1–2)
N particle count (April 25 2006) 9–17.
[7] J. Li, and L. Wang, Concentration distributions during mass transfer in circulating
R Reaction rate, mol/s fluidized beds. 7th ICCFB 2002, Niagra Falls.
Re Reynolds number [8] D. Subbarao, S. Gambhir, Gas particle mass transfer in risers. 7th ICCFB 2002,
Sc Schmidt number Niagra Falls.
[9] O. Bolland, Describing mass transfer in circulating fluidized beds by ozone
Sh Sherwood number decomposition. REPORT 1998:02.
t time, s [10] J. Kalil, J.C. Chu, W.A. Wetteroth, Mass transfer in a fluidized bed, CEP 49. No. 3
U Velocity, m/s (1953) 141.
[11] W. Resnick, R.R. White, Mass transfer in systems of gas and fluidized solids, CEP 45,
x mole fraction No 3. (1949) 377.
[12] R.H. Venderbosch, W. Prins, W.P.M. van Swaaij, Mass transfer and influence of the
local catalyst activity on the conversion in a riser reactor, Can Journal of Chemical
Greek Engineer vol. 77 (April 1999) 262.
[13] D.J. Gunn, Transfer of heat or mass to particles in fixed and fluidized beds,
ɛ Voidage International Journal of Heat and Mass Transfer 21 (1978) 467–476.
μ Viscosity, kg/m s2 [14] R. Zenenhoven, M. Jarvinen, CFB reactors, CFD and particle/turbulence interac-
ρ Density, kg/m3 tions. 4th ICMF 2001, New Orleans.
[15] R. Higbie, The rate of absorption of a pure gas into a still liquid during short periods
of exposure, Transactions of the A.I.Ch.E. 31 (1935) 365–387.
[16] P.V. Danckwerts, Significance of liquid-film coefficients in gas absorption, Industrial
Subscripts and Engineering Chemistry 43 (1951) 1460–1467.
[17] H.L. Toor, J.M. Machello, Film-penetration model for mass and heat transfer, AICHE
0 initial
Journal vol. 4 (Issue 1) (1958) 97–101.
c chemical [18] S. Banerjee, D.S. Scott, E. Rhodes, Mass transfer to falling wavy liquid films in
d mass transfer turbulent flow, I & E Ch Fundamentals vol. 7 (Issue 1) (1968) 22–27.
g gas [19] Bird, Stewart and Lightfoot, Transport Phenomena, John Wiley and Sons, New York,
2002.
i index [20] R.W. Breault, C.J. Ludlow, P.C. Yue, Cluster particle number and granular
m index temperature for cork particles at the wall in the riser of a CFB, Powder Technology
n index vol. 149 (Issues 2–3) (January 3 2005) 68–77.
[21] C. Guenther, R.W. Breault, Wavelet analysis to characterize cluster dynamics in a
p particle circulating fluidized bed, Powder Technology vol. 173 (Issue 3) (April 30 2007) 163–173.
s solid [22] M. Syamlal, MFIX document numerical technique, Department of Energy Report
sl slip DOE/MC31346-5824, Jan. 1998.

S-ar putea să vă placă și