You are on page 1of 10

QMOM-CFD MODEL DEVELOPMENT FOR AN IDEALISED PIPE

GIBBSITE PRECIPITATOR
1,3 1 2
Kieran Hutton , Darrin Stephens , Iztok Livk
1
CSIRO Mathematics Informatics and Statistics, Clayton, VIC 3168, Australia
email: kieran.hutton@csiro.au
2
CSIRO Light Metals Flagship (Process Science and Engineering)/Parker Centre,
Waterford, Perth, WA 6152, Australia
3
CSIRO Light Metals Flagship / Parker Centre, Clayton, VIC 3168, Australia

ABSTRACT

The Quadrature Method of Moments (QMOM) Population Balance technique was


implemented in a CFD 2-D axisymmetric plug-flow pipe reactor for gibbsite
precipitation to simulate the evolution of the moments of the Crystal Size Distribution
(CSD). Incorporated into the model are the effects of nucleation, growth and
agglomeration on the evolution of moments, as well as their effects on the overall mass
balance. ANSYS CFX 12.1 CFD modelling platform was used. Simulations, predicting
the CSD’s moments’ evolution along the pipe, were run under the assumption of
constant shear rate and temperature. CFD results are compared to those obtained by a
similar gibbsite precipitation model coded in MATLAB. Results from the two models,
in terms of the average particle diameters, solids and solution mass fractions, and
moments, match very closely, which confirms consistency of the 1-D CFD model
developed here for a plug-flow gibbsite precipitator.

Keywords: Computational Fluid Dynamics, Quadrature Method of Moments, Gibbsite,


Precipitation

INTRODUCTION

The precipitation stage of the Bayer Process is of key importance in the extraction of
alumina from bauxite ore. This stage consumes approximately 80% of the processing
time. As a result of this, maximising the efficiency of alumina precipitation may present
many benefits both from an economic and environmental perspective. Due to its
complex nature, it was deemed necessary to develop a CFD model with the end goal of
accurately describing the dynamics of the precipitation process for a real physical
geometry. In order to achieve this goal a simplified 1-D CFD model was developed
which incorporated the effects of molecular growth, nucleation, agglomeration and (if
required) breakage. This simplified model was developed, firstly to ensure that the
implementation of the algorithm was done correctly and also to allow for relatively
quick debugging simulations to be performed rather than debugging on a full scale
precipitator geometry.

Typically, Population Balance Modelling (PBM) involves the division of a system’s


particle size distribution into many size classes and solving the Population Balance
Equation (PBE) for each size class. The PBE may be thought of as representing the rate
Hutton et al.

of change of the number of particles of a given size due to the effects of aggregation and
breakage and this may also be extended to include the effects of molecular growth and
nucleation.
A dynamic PBE for precipitation systems with particle agglomeration was first
proposed by Smoluchowski (Smoluchowski, 1917). This equation describes the rate of
change of the number of particles of diameter, L, due to aggregation and breakage only.

a
( ) 
1
k 
 L3
− λ 3 3
, λ  ∞
(
 n L3 − λ3 ) n(λ )dλ − n(L )∫ k (L, λ )n(λ )dλ
2 L 1
dn( L) L 
= ∫
a

3

( ) (1)
2
dt 2 0
L3 − λ3 3   0


+ ∫ β (L λ )k b (λ )n(λ )dv − k b (L )n(L )
L

where L and λ are the particle diameters (m) and n(L), ka, β, kb are the number density
(# m-3), agglomeration kernel (-), fragment distribution function (-) and breakage kernel
(-) respectively.

This equation may also be expanded to account for the effects of molecular growth and
nucleation by adding the following terms,

dn( L) dn( L) (2)


=G + δ ( L − Lcrit ) BU
dt dL

where G, δ, Lcrit and Bu are the linear growth rate (m s-1), dirac delta function, critical
particle size (m) and nucleation source term (# m-3 s-1) respectively.

Several methods exist which attempt to reduce the required discretisation and therefore
reduce the computational cost of simulations (Batterham et al., 1981; Hounslow et al.,
1988; Marchal et al., 1988; Litster et al., 1995). For example the work of Batterham
introduced the geometric size discretisation where each size class contains particles of
mass twice the value of those in the preceding class (Batterham et al., 1981).

This method results in a significant reduction in the number of classes compared to


constant discretisation. The application of this method to computational fluid dynamics,
despite the reduction in classes, still results in a prohibitively large number of size
classes which must be transported in the CFD domain.

QUADRATURE METHOD OF MOMENTS (QMOM)

To combat the problem of a large number of transport equations, the quadrature method
of moments provides an alternative solution. The Quadrature Method of Moments for
precipitation and other such systems involving discrete phase population balances was
first conceived by McGraw (McGraw, 1996). This method introduces a way in which
the rate equations for the precipitation dynamics may be closed and significantly
reduces the computational effort required to accurately describe the dynamics of
precipitation. A typical discretised PB simulation may require 40 individual classes,

Chemeca 2010 Conference, Hilton Adelaide, South Australia, Australia


26-29 September 2010
2
Hutton et al.

introducing 40 new transport equations, whereas the QMOM approach requires a mere
6 additional transport equations.
Integration of Equations 1 and 2 above multiplied by the length coordinate raised to an
integer power leads to the calculation of the moment form of the rate equation. The
moments of a distribution are defined as follows,


m k = ∫ Lk n(L )dL
(3)
0

Equation 3, combined with equations 1 and 2, when expanded, leads to the moment
form of the rate equation,

∞ ∞

( 
) ( ) d (L )
k
1
 1 1
= ∫ n(λ )∫ k a  L3 − λ3 3 , λ  L3 n L3 − λ3
dmk 1 3 3
− λ3 3 dλ
dt 20 0    
∞ ∞ ∞ ∞
− ∫ Lk n(L )∫ k a (L, λ )n(λ )dλdL + ∫ Lk ∫ β (L λ )k b (λ )n(λ )dλdL
(4)
0 0 0 0
∞ ∞
− ∫ Lk k b (L )n(L )dL + ∫ kLk −1Gn(L )dL + Lkcrit BU
0 0

The calculation of the rate of change of moments in this form is very computationally
expensive due to the presence of the double integrals which cannot be performed
analytically if non-constant kernels are employed. An alternative solution to this
problem is to use a quadrature approximation to the integrals. This technique
approximates the CSD with N particle diameters (abscissas, Li) and corresponding
weighting factors (ωi), essentially representing the PSD with an N-bar bar chart. The
resulting quadrature summation is as follows,

N
mk ≈ ∑ ωi Lki (5)
i =1

The resulting dynamic equation reduces to


( )
= ∑ ω i ∑ ω j L3i + L3j 3 ψ 10 k a (Li , L j ) − ∑ Lki ω i ∑ψ 10 k a (Li , L j )ω j
k
dmk 1 N N N N

dt 2 i =1 j =1 i =1 i= j
(6)
+ ∑ k (Li )β (Li , L j )ω i − ∑ L k (Li )ω i + ∑ kL Gω i + L BU
N N N
a k a k −1 k
i crit
i =1 i =1 i −1

The additional term, ψ10, is an agglomeration efficiency term which accounts for the
fact that not every particle-particle collision results in the successful formation of an
aggregate
Three nodes (N=3) are used in these simulations as this results in acceptable accuracy
for the computational effort required (Marchisio et al., 2002). The result of this
reduction to summation terms is a significant reduction in the computational effort
required for population balance simulations.

Chemeca 2010 Conference, Hilton Adelaide, South Australia, Australia


26-29 September 2010
3
Hutton et al.

MODELLING STRATEGY AND KINETICS

Additional Variables and Mass Fractions


Suspended solids was added as an additional mass fraction. The inter-phase slip was
determined using the algebraic slip model or drift-flux method. A constant solid phase
slip diameter was employed here resulting in 1-way coupling of the fluid (continuous)
and solids (discrete) phases. The ability to employ a variable slip diameter, such as an
average particle diameter has been implemented allowing for future 2-way coupling of
precipitation and hydrodynamics.

The six moments required to perform a 3-node QMOM simulation were attached to and
transported by the solid phase.

Dissolved alumina was added as an additional mass fraction and carried with the fluid
phase. The dissolved alumina mass fraction was coupled with the solid phase mass
fraction, which accounts for inter-phase mass transport. The volumetric rate of mass
transfer from solution to solid is equal to the rate of change of the 3rd moment.

Caustic was added as an additional variable attached to the fluid phase.

Model Equations
Supersaturation is the driving force for precipitation and was determined by evaluating
the saturation concentration of alumina under the local conditions in the pipe. The
saturation concentration was determined using a semi-empirical correlation, which
accounts for the effects of chemical composition and temperature of the solution
(Rosenberg and Healy, 1996),

0.96197C
A* = (7)
1+ J

where A* and C are the saturation concentration (kg m-3) and caustic concentration (kg
m-3) respectively and parameter J is a function of chemical composition and
temperature, the details of which can be found in Rosenberg and Healy (1996).

A size-independent linear growth model was implemented using the dissolved alumina
and saturation concentration described above (White and Bateman, 1988),

 − ∆EG  ( A − A *)
2
G = kG exp  2.5 (8)
 RT  C

where G, kG, ∆EG, R, T, A and C are the linear growth rate (µm hr-1), pre-exponential
constant (70,700 J mol-1), activation energy (7.4x1012 J mol-1 K-1), universal gas
constant (8.314 J mol-1 K-1), temperature (K), dissolved alumina concentration (kg m-3
solution) and caustic concentration (kg Na2CO3 m-3 solution) respectively.

A secondary nucleation model was implemented using the following empirical


correlation (Li et al., 2003),

Chemeca 2010 Conference, Hilton Adelaide, South Australia, Australia


26-29 September 2010
4
Hutton et al.

BU = K N σ 2 γ M (9)

where KN, σ, γ and M are the rate constant, supersaturation ratio (-), fluid shear rate (s-1)
and suspended solids concentration (kg m-3 mixture) respectively.

The standard turbulent hydrodynamic kernel (Smoluchowski, 1917) was the collision
rate constant of choice here as the flow in precipitation systems is known to be in the
turbulent regime,

ε
k a = 1.29 (Li + L j ) (10)
ν

where ε and ν are the turbulent energy dissipation rate (m2 s-3) and kinematic viscosity
respectively (m2 s-1).

Agglomeration efficiency was used to incorporate the sticking efficiency of particle-


particle collisions (Ilievski and Livk, 2006). This correlation takes the form:

0.00042
ψ 10 =
−19 1.5 
1 + 6.25 × 10 η
( )
 γ 2 / U 2.25 
tip 
(12)
 (G / L )3 
 10 

where Utip, η and L10 are the impeller tip velocity (ms-1, a quantity used in the
determination of the efficiency term), kinematic viscosity (m2 s-1) and the number
averaged particle diameter (m) respectively.

At present the crystal breakage is not being accounted for in the model as it is believed
that breakage is not a significant factor under these operating conditions (Ilievski and
Livk, 2006). However, breakage has been coded into the model to allow for future
investigations.

Moment Scaling
As the code was being tested it was noted that significant instabilities were present in
the simulations. It is believed that this is as a result of the magnitude of the moments
and their source terms across the set. The typical magnitude range of number based
moments was about 22-23 orders of magnitude from ~1013 to ~10-10. These moments
were calculated using meters (m) as the length coordinate. The presence of extremely
small and extremely large numbers resulted in significant instabilities in the CFX linear
solver.
To combat this problem it was proposed that the moments be scaled and these scaled
moments be transported by the CFX solver. The scaling resulted in the moments being
calculated on a micron basis which reduced the range from the aforementioned 22-23
orders of magnitude to a mere 7-8 orders. This stabilised the CFD simulations and
subsequently resulted in shorter simulation times.

Chemeca 2010 Conference, Hilton Adelaide, South Australia, Australia


26-29 September 2010
5
Hutton et al.

It must be noted that the QMOM-PB algorithm is still run on a meter length scale and
then converted to the micron based moment scale using a scaling factor of 106k-18
(where k is the moment number).

1-D GIBBSITE PRECIPITATOR

Fig 1. Gibbsite Precipitator Domain Geometry.


Fig 1 shows the wedge used as the 2D axisymmetric domain. Results for output were
chosen along the centreline of this virtual pipe. In order to test the model it was decided
to compare the results to those obtained from a MATLAB QMOM-PB model. The
results of these simulations were not expected to be truly representative of real
precipitation systems, although they are based on an established foundation of empirical
correlations, but were used to verify the correct implementation of the QMOM-PB code
in CFX.
The test cases were all run at the same initial conditions and the three phenomena
(agglomeration, growth and nucleation) were tested individually and in combination.
The initial conditions used in the test simulations were as follows:
Tab 1. Initial Conditions Employed in Test Simulations.
Quantity Value Units
Dissolved Alumina, A 170 kg Al(OH)3 m-3 Fluid
Caustic, C 240 kg Na2CO3 m-3 Fluid
Solids Loading, M 250 kg Solids m-3 Mixture
Temperature, T 335.15 K
Shear Rate, γ 500 s-1
Impeller Tip Speed, Utip 2.12 (arbitrary) ms-1
Dynamic Viscosity, η 4x10-3 Pa s

The seed distribution chosen for the test cases was a Gaussian distribution with a mean
and standard deviation of 30µm and 5µm respectively. The details of these initial
moments are given below in Tab 2.
Tab 2. Raw and Scaled Inital Distribution's Moments.
Moment Number Raw(mk) Scaled (µmk)
m0 6.25105x10+12 6.25105x10-06
m1 1.92741x10+08 1.92741x10-04

Chemeca 2010 Conference, Hilton Adelaide, South Australia, Australia


26-29 September 2010
6
Hutton et al.

m2 6.09479x10+03 6.09479x10-03
m3 1.97299x10-01 1.97299x10-01
m4 6.52847x10-06 6.52847x10+00
m5 2.20517x10-10 2.20517x10+02

RESULTS & DISCUSSION

Each of the figures presented are labelled with a prefix code AGN, which is a logic code
indicating which phenomena’s source terms are switched on and off. For example,
AGN_010 has molecular growth as the only source term with all others set to zero.

0th Moment, m0 – Total Particle Count


6.60E+12

6.55E+12

6.50E+12
0 Moment, m0 (# / m Mixture)

6.45E+12
3

6.40E+12

6.35E+12

6.30E+12
th

6.25E+12

6.20E+12

6.15E+12
0 2000 4000 6000 8000 10000 12000 14000
Time (s)
AGN_001 CFD AGN_001 Matlab AGN_010 CFD AGN_010 Matlab
AGN_110 CFD AGN_110 Matlab AGN_111 CFD AGN_111 Matlab

Fig 2. 0th Moment Evolution.


Figure 2 shows the evolution of the 0th moment of the CSD with various source terms
switched on or off. The case of pure nucleation (AGN_001) can be seen to have the
most dramatic impact on m0. This is as a result of there being only a very slight drop in
the supersaturation ratio over the course of the simulation (see Fig. 4). Nucleation
continues at very close to its initial rate. The case of crystal growth alone (AGN_010)
has no impact on m0 as no new particles are being introduced to the system. The case of
agglomeration and growth (AGN_110) has the effect of reducing the total particle
count. This is effected by agglomeration alone as growth has no impact on m0. With the
three source terms turned on, the effect is an initial decrease in count due to
agglomeration being the dominant mechanism followed by an increase in count as
nucleation becomes dominant due to a reduced sticking efficiency brought about by a
reduction in supersaturation ratio.
In all four cases the CFD and MATLAB simulated outputs are almost identical, which
can be seen by the coincidence of output plots.

Chemeca 2010 Conference, Hilton Adelaide, South Australia, Australia


26-29 September 2010
7
Hutton et al.

3rd Moment, m3 – Total Particle Volume


0.27

0.26
3 Moment, m3 (m / m Mixture)

0.25

0.24
3
3

0.23

0.22
rd

0.21

0.2

0.19
0 2000 4000 6000 8000 10000 12000 14000
Time (s)
AGN_001 CFD AGN_001 Matlab AGN_010 CFD AGN_010 Matlab
AGN_110 CFD AGN_110 Matlab AGN_111 CFD AGN_111 Matlab

Fig 3. 3rd Moment Evolution.


The evolution of the 3rd moment is hardly affected by nucleation alone. This is because
the total volume of nuclei is very small relative to the total volume of existing particles.
Each of the other three cases examined follow almost exactly the same curve, with m3
increasing over time. Crystal growth is the dominant factor here increasing the total
particle volume. Agglomeration has no direct impact on m3 as it does not affect the total
volume of particles. In each of these cases the CFD and MATLAB outputs are almost
identical.

Alumina Supersaturation Ratio, σ

1.79

1.59
Alumina Supersaturation Ratio, σ (-)

1.39

1.19

0.99

0.79

0.59

0.39

0.19
0 2000 4000 6000 8000 10000 12000 14000
Time (s)
AGN_001 CFD AGN_001 Matlab AGN_010 CFD AGN_010 Matlab
AGN_110 CFD AGN_110 Matlab AGN_111 CFD AGN_111 Matlab

Fig 4. Alumina Supersaturation Ratio Evolution.

Chemeca 2010 Conference, Hilton Adelaide, South Australia, Australia


26-29 September 2010
8
Hutton et al.

Molecular growth has the largest impact on the evolution of the alumina supersaturation
ratio. Growth has the effect of consuming solute from solution and transferring it to the
solid phase by surface deposition. Nucleation has negligible impact on supersaturation.
Agglomeration has no noticeable effect on supersaturation.

Volume Averaged Particle Diameter, d43

3.65E-05

3.60E-05
Volume Averaged Particle Diameter (m)

3.55E-05

3.50E-05

3.45E-05

3.40E-05

3.35E-05

3.30E-05

3.25E-05
0 2000 4000 6000 8000 10000 12000 14000
Time (s)
AGN_001 CFD AGN_001 Matlab AGN_010 CFD AGN_010 Matlab
AGN_110 CFD AGN_110 Matlab AGN_111 CFD AGN_111 Matlab

Fig 5. Volume Averaged Particle Diameter Evolution.


The volume averaged particle diameter is defined as the ratio of the 4th and 3rd
moments. The effect of nucleation on this quantity is negligible because the volume of
nuclei is very small in comparison to the volume of existing particles. Growth alone, as
would be expected, has a significant positive impact on the average diameter. With the
addition of agglomeration, this effect is further increased resulting in a larger average
(and a decreased number). Again, the CFD and MATLAB simulation outputs yield
almost identical results.
Comparison of other process parameters (not shown here), output by both models, such
as the evolution of number averaged particle diameter, quadrature weights, quadrature
abscissas, linear growth rate and nucleation rate yield near identical results.

CONCLUSIONS

The presented results show that the development of the precipitation code for a pipe
crystalliser was successful in incorporating the effects of agglomeration, growth and
nucleation. This is evident from the fact that the output plots from the CFD and
MATLAB models are almost indistinguishable from each other.
The model, although not yet validated, shows the trends that would be expected from a
gibbsite precipitating system operating under the investigated conditions.
The next stage in the code development is the inclusion of 2-way coupling of the
discrete phase with the hydrodynamics and the subsequent extension of this method to
full 3-dimensional flow in a stirred tank and the validation of the model and/or

Chemeca 2010 Conference, Hilton Adelaide, South Australia, Australia


26-29 September 2010
9
Hutton et al.

modification of kinetic parameters to enable the simulation of real gibbsite precipitation


systems.

REFERENCES

Batterham, R.J., Hall, J.S., Barton, G., 1981. “Pelletizing kinetics and simulation of full scale
balling circuits”. Proceedings of the 3rd International Symposium on Agglomeration, Nurnberg,
A136-150.

Hounslow, M.J., Ryall, R.I., Marshall, V.R., 1988. “A discretized population balance for
nucleation, growth and aggregation”. AIChE J. 34, 11, p. 1821–1832.

Ilievski, D. and Livk, I., 2006. “An Agglomeration Efficiency Model for Gibbsite Precipitation
in a Turbulently Stirred Vessel”. Chemical Engineering Science, 61(6), p. 2010-2022.

Li, T.S., Livk, I., Lane, G., Ilievski, D., 2003. “Dynamic Compartment Models of Uniformly-
Mixed and Inhomogeneously-Mixed Gibbsite Crystallisers”. Chem Eng Technol 26, 3, p. 369-
376.

Litster, J.D., Smit, D.J., Hounslow, M.J., 1995. “Adjustable discretised population balance for
growth and aggregation”. AIChe Journal, Vol. 41. No. 3, p. 591-603.

Marchal, P., David, R., Klein, J.P., Villermaux, J., 1988. “Crystallization and precipitation
engineering - I. An efficient method for solving population balance in crystallization with
agglomeration”, Chem. Engng Sci. 43, p. 59–67.

Marchisio, D.L., Vigil, R.D., Fox, R.O., 2002 “Quadrature Method of Moments for
Aggregation-Breakage Processes”. Journal of Colloid and Interface Science, 258, p. 322-334.

McGraw, R., 1996. “Description of Aerosol Dynamics by the Quadrature Method of Moments”,
Aerosol Science and Technology, 27(2), p. 255-265.

Rosenberg, SP. and Healy, S.J., 1996. “A thermodynamic model for gibbsite solubility in Bayer
liquors”, Proceedings of 4th International Alumina Quality Workshop, Darwin, NT.

Smoluchowski, M.Z., 1917. “Versuch Einer Mathematischen Theorie Der Koagulationskinetik


Kolloider Losunger“. Zeitschrift fur Physikalische Chemie, 92, p. 129-142.

White, E.T. and Bateman, S.H., 1988. “Effect of caustic concentration on the growth rate of
Al(OH)3 particles”. Light Metals p. 157-162.

Chemeca 2010 Conference, Hilton Adelaide, South Australia, Australia


26-29 September 2010
10

You might also like