You are on page 1of 96

David G.

Dixon Mixing and Mass Transfer in Hydrometallurgy

Mixing and Mass Transfer


in Hydrometallurgy

David G. Dixon
University of British Columbia
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Mass Transfer and Chemical Reaction


in Leaching Systems

Leaching operations involve the following steps:


a) Absorption of gases into solution
b) Reaction between the gases and redox catalysts
c) Diffusion of reactants to the mineral surface
d) Adsorption/reaction/desorption at the mineral surface
e) Diffusion of products away from mineral surface
f) Reaction between redox catalysts and product solutes
g) Desorption of product solutes into the gas phase
h) Reaction in the gas phase
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Transfer of Dissolved Oxygen

O2
Gas bubble
FeS2
O2

Fe3+, SO42-

The mechanism of pyrite autoclaving (a, c, d, e)


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Leaching with a Surrogate Oxidant

O2
3+
Fe
Gas bubble
2+ + 3+ ZnS
O2 + 4 Fe + 4 H = 4 Fe + 2 H2O
O2
2+
Fe

2+
Zn

The mechanism of zinc pressure leaching (a, b, c, d, e)


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Heterogeneous Reduction of Oxygen

+ + 2+
O2 + 4 Cu + 4 H = 4 Cu + 2 H2O
2+
Cu

Gas bubble
2+ n+ + (n+1)+
Cu +S = Cu + S CuFeS2
O2 or Ni3S2
Cu+ Sn+

+
Cu
3+ 2+
Fe or Ni

The mechanism of ammoniacal leaching of sulfides (a, b, c, d, e, f)


(Anaconda-Arbiter, Sherritt Gordon)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Gas-Phase Conversion of Oxygen

NO

Gas bubble
+ +
NO2 + NO + 2 H = 2 NO + H2O FeAsS
2 NO + O2
= 2 NO2
NO2 NO+

Fe3+, AsO43-, SO42-

The mechanism of the NITROX process (a, b, c, d, e, g, h)


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Applications of Fluid Mixing

Physical Processing Application Class Chemical Processing


Suspension Liquid-Solid Dissolution
Dispersion Liquid-Gas Absorption
Emulsification Immiscible Liquids Extraction
Blending Miscible Liquids Reaction
Pumping Fluid Motion Heat Transfer
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Axial-Flow Impellers
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Radial-Flow Impellers
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Axial-Flow Impeller, Open Tank

No baffles: vortexing Baffles: no vortexing


Single circulation loop
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Radial-Flow Impeller, Open Tank

No baffles: vortexing Baffles: no vortexing


Double circulation loop
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Radial-Flow Impeller, Closed Tank

No vortexing, no baffles required


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Reference Tank Geometry

Important for comparing performance of impellers


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Mixing Dynamics

Application Impeller Types


Heat Transfer Hydrofoils
FLOW Blending Propellers
Solids Suspension Axial Flow Impellers
Gas-Liquid Dispersion Radial Flow Impellers
SHEAR Liquid-Liquid Dispersion Flat-Blade Turbines
Emulsification Disc Turbines
De-agglomeration Homogenizers
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Mixing Dynamics
By dimensional analysis:

P  N 3D5 ρ
Q  N D3
H  N 2D2
hence P  H Q ρ

where:
P = power applied to the impeller (W)
Q = pumping capacity of the impeller (m3/s)
H = velocity head (or shear) developed by the impeller (m2/s2)
N = impeller speed (1/s)
D = impeller diameter (m)
ρ = fluid density (kg/m3)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Flow vs Shear
At constant power:

Q 1
= D8 / 3 = 8 / 5
H N

Hence, for a constant amount of power applied

• increasing the impeller diameter increases the flow capacity of the


impeller dramatically relative to its shear generating capacity

• increasing the rotational speed does just the opposite, albeit somewhat
less dramatically.

For shear, small impellers run at high rotational speeds are preferred.
For flow, large impellers run at low rotational speeds are preferred.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Flow vs Shear

High fluid shear generated primarily at impeller tip

Efficient dispersion requires the discrete phase pass through the


impeller zone, which requires flow pumping capacity and baffling
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Flow vs Shear
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Power number vs Reynolds number


The dynamics of various impeller designs are described in terms of a
dimensionless "Power number” (or “Newton number”) which is defined:

P
N Po  N Ne  3 5
N D ρ

This number represents a "drag coefficient" or "friction factor" for impellors, and
can therefore be correlated with the impeller Reynolds number:

D2 N ρ D2 N
N Re  
μ ν

where:
μ = kinematic viscosity (kg/m/s)
ν = dynamic viscosity (m2/s)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Power number vs Reynolds number


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Constant Power numbers at high Reynolds numbers


Impeller type and configuration Power number
3-blade propeller, pitch = D 0.32
3-blade propeller, pitch = 2D 1.00
Shrouded turbine with stator ring 1.12
2-blade flat paddles, D/W = 8 1.15
2-blade flat paddles, D/W = 6 1.60
6-blade 45 pitched-blade turbine 1.65
2-blade flat paddles, D/W = 4 2.25
4-blade flat paddles, D/W = 6 2.74
6-blade flat paddles, D/W = 6 3.82
4-blade flat-blade turbine 4.50
6-blade flat-blade turbine 6.30
8-blade flat-blade turbine 7.80
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Liquid-Solid Mass Transfer


Consider the following chemical reaction at a mineral surface:

A(aq) + b B(s)  products

The rate of liquid-solid mass transfer of dissolved species A per unit surface area
of the mineral particle is expressed:

j A  k m (C Ab  C As )

where:
jA = flux of species A to the mineral surface (mol/m2/s)
km = mass-transfer coefficient (m/s)
CAb, CAs = concentrations of species A in the bulk solution and at the mineral
surface, respectively (mol/m3)

When the reaction at the surface is fast relative to liquid-solid mass transfer, then
the mass transfer step will be rate controlling, and CAs may be taken as zero.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

The Harriott Equation


Mass transfer between a spherical solid particle and a fluid moving relative to the
particle may be described by the empirical correlation of Ranz and Marshall:

1/ 2 1/ 3
km d p  d pv p ρ   μ 
N Sh, p  2  0.6 N Re,
1/ 2
p N 1/3
Sc or  2  0.6    
DA  μ   ρ DA 

where:
NSh,p = the particle Sherwood number (analogous to the Nusselt number in
heat transfer)
NRe,p = the particle Reynolds number
NSc = the Schmidt number (analogous to the Prandtl number in heat
transfer)
dp = particle diameter (m)
DA = diffusivity of species A (m2/s)
vp = slip velocity of a particle relative to the surrounding fluid (m/s)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

The Harriott Equation


Assuming a) free settling conditions and b) that Stokes's law applies, then the slip
velocity of a particle relative to the surrounding fluid may be taken as the Stokes
settling velocity:

d p2 g ( ρ p  ρ)
vp 
18 μ
where:
ρp = solid particle density (kg/m3)

Substituting into the Ranz and Marshall correlation and solving for the mass-
transfer coefficient gives the Harriott equation:


DA   d p g ρ ( ρ p  ρ)
3

1/ 2
 μ 
1/ 3 
km = 2 + 0.6     
dp   2  
  18 μ   ρ DA  
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

The Harriott Equation


For sizes above about 100 μm, Stokes's law no longer applies, and the Harriott
equation overestimates the mass transfer coefficient. In this case, we can make
use of a well-known intermediate settling law:
1
 d 1p.6 g ( ρ p  ρ)  1.4
vp   0 .4 0 .6  @ 1.92  N Re, p
 14 ρ μ 
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

The Harriott Equation at Different Particle Densities


-3.6

-3.7

Stokes' law Intermediate law


-3.8

8
log km (m/s)

-3.9 7
6
5
4
-4.0
3

2
-4.1

-4.2
1
 p   = 0

-4.3
1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0
log d p (mm)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

The Effect of Agitation


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Solids Suspension – The Zwietering Equation


For achieving complete off-bottom conditions in the absence of gas sparging, the
correlation of Zwietering (1958) is almost universally employed:

 T T  ν d p [ g ( ρ p  ρ) / ρ] X
0 .1 0 . 2 0.45 0.13

N js = S  , 
D C D 0.85
where:
Njs = impeller speed to just suspend solids (s–1)
S = dimensionless function of system geometry
T = tank diameter (m)
C = impeller clearance from tank bottom (m)
dp = Sauter mean particle size (m)
X = pulp density (%)

Note that this equation only describes the critical impeller speed to lift particles
off the tank bottom. It does not describe complete particle dispersion, which is
unnecessary for liquid-solid mass transfer as discussed above.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Solids Suspension – The Zwietering Equation


Armenante and Nagamine (1998) have quantified the S function for several single
impeller configurations:

1.40
T   C
DT * S  0.99   exp  2.18  (6-blade disk (Rushton) turbine)
D  T
1.20
T   C
FBT * S  1.43   exp 1.95  (6-blade flat-blade turbine)
D  T
0.83
T   C
PBT ** S  2.28   exp  0.65  (6-blade 45 pitched-blade turbine)
D  T
0.79
T   C
HE - 3** S  3.49   exp  0.66  (Chemineer HE-3 axial flow propeller)
D  T

*
Correlation only valid if lower circulation loop is suppressed
**
Pitched and axial-flow impellers pumping downward
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Solids Suspension – The Zwietering Equation


Hence, for the reference tank geometry shown above:

D C 1
= =
T T 3

the following coefficients apply to the Zwietering correlation:

DT * S  9.53
FBT * S  10.24
PBT ** S  7.05
HE - 3** S  10.36

Hence, the critical impeller speed is not that sensitive to impeller type under
ungassed conditions, although it is very sensitive to impeller placement above the
tank bottom (as evidenced by the fact that S is an exponential function of C/T),
and especially so for radial-flow impellers.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Validation of Exponential S-functions

DT FBT

PBT HE-3
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Solids Suspension – The Zwietering Equation


Also, since, according to Zwietering, for a given fluid and solid,

N js  D 0.85  constant

then since, under turbulent mixing conditions,

P  N 3 D 5 and given that V  D3

then it follows that

P
 D 0.55
V

Hence, the required power per unit volume (specific power) for solids suspension
decreases as the dimensions of the tank increase.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Segregation of Just-Suspended Solids

Coarsest solids remain near tank bottom


Solids distribution depends on location of draw-off
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

EXAMPLE 1 – Impeller power requirement for


off-bottom suspension of pitchblende.

Determine the impeller power required to suspend pitchblende (U3O8) particles


with a Sauter mean diameter of 100 μm in water at a pulp density of 20% in a
tank 3 m in diameter. Determine the power as a function of impeller diameter D
and impeller clearance C for both a standard six-blade disk (Rushton)
turbine (DT) and a standard downward-pumping six-blade 45 pitched-blade
turbine (PBT).

Solution. We will solve Zwietering’s correlation using the geometric functions


of Armanente and Nagamine for DT and PBT, given the following:

ν = 10–6 m2/s
dp = 100 μm = 10–4 m
g = 9.81 m/s2
ρp = 8,380 kg/m3
X = 20% (by mass)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

EXAMPLE 1 – (continued)

To calculate the suspension power requirement, we will utilize the Power number
equation assuming turbulent flow:

Pjs  N Po ( N 3js D 5 ρ)

taking power numbers from the previous table:

NPo (DT) = 6.30


NPo (PBT) = 1.65

In order to achieve the most conservative estimate, we will employ the slurry
density in the Power number equation:
100 100
ρ slurry    1,214 kg/m 3
100  X X 80 20
 
ρ ρ p 1,000 8,380
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

"Just suspend" rotational speed: DT


D/T =
5.5
0.25
5.0

4.5

4.0

3.5 0.30
Njs (Hz)

3.0

2.5 0.35

2.0
0.40
1.5
0.45
1.0 0.50

0.5
0.03 0.06 0.09 0.12 0.15 0.18
C/T
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

"Just suspend" rotational speed: PBT

5.0
D/T =
0.25
4.5

4.0

3.5
0.30
Njs (Hz)

3.0

0.35
2.5

0.40
2.0
0.45
1.5 0.50

1.0
0.1 0.15 0.2 0.25 0.3 0.35
C/T
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Power requirement: DT

300
D/T =
0.25
250

200
0.30
Pjs (kW)

150 0.35

0.40
100
0.45
0.50

50

0
0.03 0.06 0.09 0.12 0.15 0.18
C/T
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Power requirement: PBT

49

46

43

40
Pjs (kW)

37
D/T = 0.25 D/T = 0.50

34

31

28
0.10 0.15 0.20 0.25 0.30 0.35
C/T
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

The Effect of Gas Sparging

Impeller flooding = catastrophic loss of solids suspension


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

The Effect of Gas Sparging


Several modifications to Zwietering’s equation have been proposed in the
presence of gas sparging. One of the more recent ones which is very easy to use
is as follows [Nienow and Bujalski (2002)]:

N js ( g )  N js (0.83  0.31 Q g )

where
Njs(g) = impeller speed to just suspend solids with gas sparging (s–1)
Qg = gas flowrate (volumes of gas per volume of tank per minute or vvm)

The suspending ability is initially enhanced at low gas rates due to the increased
in fluid flow induced by the gas, but is diminished once the gas rate exceeds 0.55
vvm due to the onset of impeller flooding.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Application I: Dissolution of Salts


Dissolution of a soluble salt in water is purely a function of liquid-solid mass
transfer. The rate of dissolution is governed by the rate at which the product ions
can diffuse away from the salt crystals. For example, to calculate the rate of
dissolution of table salt:

NaCl(s)  Na+ + Cl–

requires solving the following set of algebraic equations:

 s
j Na  = (k m ) Na  C Na  C
b
Na 
 
j Cl = (k m ) Cl C Cls   C Clb  
j Na   j Cl s
C Na   C  = K sp , NaCl
s
Cl

where the km values would be calculated from the Harriott equation.


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Dissolution of Salts
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Application II: Acid Dissolution of Simple Oxides


In this system, the rate is limited by the diffusion of protons to the mineral
surface, as well as of metal ions away from the surface. For example, in order to
determine the rate of the acid dissolution of zinc calcine:

ZnO(s) + 2 H+  Zn2+ + H2O

the following equations must be solved simultaneously:


j Zn 2   (k m ) Zn 2  C Zn
s
2  C
b
Zn 2 
 
j H   ( k m ) H  C Hb   C Hs  
s
C Zn 2
2 j Zn 2   j H   K eq, ZnS
C 
s
H
2
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Acid Dissolution of Simple Oxides


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Application III: Mass Transfer of Dilute Reagents


A classic case of the rate of leaching limited by the diffusion of a dilute reagent
to the mineral surface is the carbonate leaching of uranium oxide, or pitchblende.
Here, the reaction is:

U3O8 + 6 HCO3– + 3 CO32– + 1/2 O2  3 UO2(CO3)34– + 3 H2O

In this case, the rate of leaching is given simply by:

1 dN U3O8
 rU3O8    2 k m C Ob 2
S U 3O 8 dt
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Mass Transfer of Dilute Reagents

Carbonate leaching of pitchblende


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

EXAMPLE 2 – Mass transfer of a dilute reagent during leaching.

Estimate how long it would take to dissolve a 100-μm pitchblende particle in


sodium carbonate solution in equilibrium with air at 60°C.

Solution. Recasting the rate law for pitchblende dissolution in terms of particle
diameter thus:

1 dN U3O8 ρ U 3O 8 dVU3O8 ρ U 3O8 d dp


rU 3O8      2 k m C Ob 2
S U 3O 8 dt M U 3O 8 S U 3O 8 dt 2 M U 3O8 dt

then we can define the particle shrinkage rate in terms of the Harriott equation:

d dp 4 M U 3O8 k m C Ob 2

dt ρ U 3O 8

4 M U 3 O 8 DO 2 C b  1/ 2
 d g ρ ( ρ U 3O 8  ρ )   μ
3

1/ 3


O2
2 + 0.6  p    
ρ U 3O 8 d p   18 μ 2   ρ DO  
    2  
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

EXAMPLE 2 – (continued)

Defining dimensionless particle size thus:

dp
ξ
d p ,0

then the rate equation becomes

4 M U 3O8 DO 2 C b   d p , 0 g ρ ( ρ U 3O 8  ρ )
3

1/ 2
 μ 
1/ 3

dξ O2 2
 + 0.6     
 ξ 1/ 2

dt ρ U 3O8 d p2, 0 ξ  18 μ 2   ρ DO  
    2  
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

EXAMPLE 2 – (continued)

Given the following:

ρ U 3O8= 8,380 kg/m3


M U 3O8 = 842.08 kg/kmol
DO 2 = 2.75  10–9 m2/s
C Ob 2= 1.83  10–4 kmol/m3 (air at 60°C, 1 atm, effect of Na2CO3 ignored)
μ = 10–3 Pa-s
ρ = 1,000 kg/m3
dp 0 = 100 μm = 10–4 m

then the rate equation is reduced to the following:

dξ 2 
 2.0  10 5   8.6 ξ 
dt ξ 

By numerical integration, the time required to fully dissolved the 100-μm U3O8
particle would be 4,200 s or 70 min. Ans.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Gas-Liquid Mass Transfer: Physical Adsorption


When a gas dissolves into a liquid without reacting, the rate of absorption is
expressed:

  
R  k g a p Ab  p A*  k l a C A*  C Ab 
where:
R = volumetric absorption rate (kmol/m3/s)
kg = gas-side mass transfer coefficient (kmol/m2/s/Pa)
kl = liquid-side mass transfer coefficient (m/s)
a = volume-specific gas-liquid interfacial area (1/m)
pAb, pA* = partial pressure of the dissolving gas in the gas bulk and at the
interface, respectively (Pa)
CAb, CA* = concentration of the dissolved gas in the liquid bulk and at the
interface, respectively (kmol/m3)

The gas-side resistance to mass transfer is nearly always negligible.


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Liquid-Side Mass Transfer Models

Film model Surface Renewal Models


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Adsorption with Chemical Reaction – Hatta Number


Consider the following chemical reaction between a dissolved gas and an
aqueous solute:

A(g) + b B(aq)  products

whose rate is described by a simple second-order rate expression:

r  k CA CB

If steady-state conditions are assumed in the film, then material balances may be
written for the two reactants. An analytical solution of these equations is not
possible, but numerical solutions may be computed for a range of variables.
These solutions are expressed in terms of an enhancement factor E defined by:

R  E k l a C A*
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Adsorption with Chemical Reaction – Hatta Number


When greater than one, E represents the ratio of the average rate of absorption
into an agitated liquid, in the presence of a reaction, to the average rate of pure
physical absorption when the bulk liquid concentration of dissolved gas A is
zero.

E is plotted as a function of the dimensionless Hatta number:


1
N Ha = DA k C Bb
kl

the instantaneous enhancement factor:


DB C Bb
Ei  1 
b DA C A*
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Liquid-Side Concentration Profiles

Very slow reaction in Moderately slow


bulk liquid reaction in bulk liquid
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Liquid-Side Concentration Profiles

Moderately fast Fast reaction in


reaction diffusion film
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

van Krevelen-
Hoftijzer Solution:

N Ha η
E

tanh N Ha η 
Ei  E
where η 
Ei  1
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Liquid-Side Concentration Profiles

Special case:
Special case:
Fast pseudo-first-order
Instantaneous reaction
reaction
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Special Case I: Fast Pseudo-First-Order Reaction


If the concentration of component B in the bulk liquid is much greater than CA*,
then reactant B diffuses toward the surface fast enough to prevent the reaction
from causing any significant depletion there, so that CBb is kept virtually
constant.

If NHa > 3, then tanh (NHa)  1, and

E  N Ha or R  DA k C Bb a C A*

This result has two very interesting consequences:

• the rate of gas-liquid mass transfer is completely independent of the liquid-


side mass transfer coefficient

• the kinetics of the chemical reaction will appear to be ½-order in CB


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Special Case II: Instantaneous Reaction


When the rate constant is very high and the concentration of B in the bulk liquid
is low, both A and B are completely depleted at some point within the film.

The actual reaction kinetics are immaterial (and virtually unmeasurable), and the
rate of reaction is limited by the rate at which both components can diffuse
toward the “extinction plane.” In this case,

E  Ei and hence, R  E i k l a C A*

In hydrometallurgy, this condition is largely satisfied when O2 gas is absorbed


into a solution within which the cuprous form of copper is stable, including NH 3
and Cl– media at all temperatures, and SO42– media at temperatures exceeding
200°C (Kimweri, 2001).
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Modeling the Cu-NH3-O2 System

0.0014
Ammonia data, 60°C, 10 g Cu
2
(Ha*) > 128
0.0012 Model Fits
32

0.0010 16
Oxygen rate, mol/s

8
0.0008
4

2
(Ha*) = 2
0.0006

0.0004

0.0002

0.0000
0 20 40 60 80 100 120 140
Time, s
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Modeling the Cu-SO4-O2 System @ 200°C

0.0016
Sulfate data, 200°C, 10 g Cu
2
0.0014 (Ha*) > 128
Model Fits

0.0012 64
Oxygen rate, mol/s

32
0.0010
16

0.0008
8

0.0006 4

2
(Ha*) = 2
0.0004

0.0002

0.0000
0 20 40 60 80 100 120 140
Time, s
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Gas-Liquid Mixing
The mixing of a gas into a liquid depends on two distinct processes occurring in
series:

• First, the gas must get into the liquid by one of three mechanisms:

• Sparging
• Surface aeration
• Gas pumping

• Next, the gas bubbles must be broken up and dispersed throughout the
liquid volume. This is the operation which requires the high shear.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Gas Sparging Without Agitation


In the absence of agitation, a certain amount of gas dispersion is accomplished
simply by sparging. In this scenario, the gas holdup can be calculated thus:

vs
φ
vt

where:
vs = superficial velocity of gas flow in the tank (m/s)
vt = terminal rise velocity of the bubbles (m/s)

In non-electrolyte solutions, bubble diameters are typically in the range of 2 to 4


mm, and the terminal velocity may be taken as a constant value of 0.265 m/s. In
electrolyte solutions, bubble diameters are typically much smaller, and their
terminal velocities may be calculated on the basis of Stokes’ law:
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Gas Sparging Without Agitation


In non-electrolyte solutions, bubble diameters are typically in the range of 2 to 4
mm, and the terminal velocity may be taken as a constant value of 0.265 m/s. In
electrolyte solutions, bubble diameters are typically much smaller, and their
terminal velocities may be calculated on the basis of Stokes’ law:

d b2 g ( ρ  ρ g )
vt 
18 μ
where
db = bubble diameter (m)
ρg = gas density (kg/m3)

The gas-liquid interfacial area is given directly by the gas holdup as follows:

a
db
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Gas Sparging With Agitation


Even with agitation, there is a critical impeller speed below which the impeller
has virtually no effect on gas dispersion, given by:

N0 D T
1/ 4
 A B
(σ g / ρ ) D

where:
N0 = the critical impeller speed (s–1)
σ = surface tension (N/m)
A, B = impeller-specific constants

For turbine impellers: A = 1.22 and B = 1.25

For four-bladed paddles: A = 2.25 and B = 0.68


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Critical Impeller Speed for Gas Dispersion


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Flooding Loading Dispersion


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Sparging Power
The power imparted by the sparged gas on the fluid in a tank can be determined
by writing a mechanical energy balance (Bernoulli’s equation) for the gas
between location 1 (just above the sparger orifices) and location 2 (at the liquid
surface):

v 22  v12 p2
dp
 g ( z 2  z1 )   W  Lf  0
2 p1 ρ g

where
z = height (m)
p = pressure (Pa)
W = specific work (J/kg)
Lf = specific friction loss (J/kg)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Sparging Power
Ignoring friction losses, neglecting the gas velocity at the liquid surface, defining
the gas density in terms of the ideal gas law, and multiplying by the gas mass
flow rate gives the sparging power as follows:

 v12 RT p2 
Psp  m g W  m g   g ( z 2  z1 )  ln 
 2 M g p1 

where
Psp = sparging power (W)
m g = gas mass flow rate (kg/s)
R = universal gas constant (8314.3 J/kmol/K)
T = temperature (K)
Mg = gas molecular weight (kg/kmol)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Gassed vs Ungassed Impeller Power


Several investigators have attempted to correlate the effect of gassing rate on the
loss of impeller power. The correlation of Hughmark (1980) appears to be the
best for flat-blade turbines and defines the ratio of gassed to ungassed power
thus:
1/ 4 1/ 5
Pg  N V   gWV 2 / 3 
 0.10   
 4 

P0   2
 Qg   N D 

where
V = fluid volume (m3)
Qg = gas volume rate (m3/s)
W = turbine blade width (m)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

EXAMPLE 3 – Estimating gassed power in a pressure oxidation autoclave.

Estimate the ratio of gassed to ungassed power for a Rushton turbine in the first
chamber of a pressure oxidation autoclave operating at 220°C and 3200 kPa, with
a working volume of 30 m3 and a design pure oxygen flow rate of 6.7 t/h.
Assume in impeller diameter D = 1.35 m, a rotational speed N = 1.70 Hz, and an
impeller blade width W = D/5 = 0.27 m.

Solution. It is important to realize that the gas bubbles in this scenario will
consist primarily of water vapor. Hence, the “true” gassing rate which the
impeller sees will be a significantly larger than the gas volume flow rate
attributable solely to oxygen. The true gassing rate is calculated from the ideal
gas law and the vapor pressure of water as follows:

n O2 RT m O2 RT (6700)(8314.3)( 220  273.15) m3


Qg     0.271
pO 2 M O2 ( p  p H 2O(T) ) (3600)(32)(3200  2320)(1000) s
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

EXAMPLE 3 – (continued)

Using Hughmark’s correlation, the gassed power ratio is estimated as follows:

1/ 4 1/ 5
Pg  NV   gWV 2/3

 0.10    2 4 
P0 Q 
 g   N D 

1/ 4 1/ 5
 (1.70)(30)   (9.81)(0.27)(30) 2/3

 0.10    2 4
  0.45  45%
 (0.271)   (1.70) (1.35) 

Hence, according to Hughmark, the power draw of the impeller will be less than
half of what we would have calculated from the Power number equation.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Interfacial Area and Gas Holdup


At moderate impeller speeds, only the gas fed to the impeller produces the
specific interfacial area (designated a0), but at higher impeller speeds, especially
in smaller vessels, gas drawn from the space above the liquid contributes
significantly to the interfacial area a.

For disk flat-blade turbines, Calderbank (1958) has correlated gas-liquid


interfacial area with gassed impeller power and impeller Reynolds number:

0.4 1/ 2 0.7 0.3


 Pg  ρ 0.2
 vs   D N ρ   DN 
2
a  a 0  1.44     @      30,000
V  σ 0.6  vt   μ   vs 
0.7 0 .3 0.7 0.3
a  D N ρ   DN 
2
 D N ρ   DN 
2
 8.33  10 5      1.5 @      30,000
a0  μ   vs   μ   vs 
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Interfacial Area and Gas Holdup


The mean bubble diameter for disk flat-blade turbines is given by:
0.4 1/ 4
 V  σ 0.6 m  μ g 
db  K   φ  
 P  ρ 0 .2
 g  μ 
where
μg = gas viscosity (Pa-s)
K = 2.25 and m = 0.4 for electrolyte solutions
K = 1.90 and m = 0.65 for non-electrolyte solutions

In non-electrolyte solutions, the terminal bubble velocity may be taken as 0.265


m/s, according to Calderbank. However, in electrolyte solutions (which are of
the most interest to us hydrometallurgists) the terminal velocity must be
calculated from Stokes’ law, and the above set of equations must be solved
iteratively.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Interfacial Area and Gas Holdup


Finally, Calderbank also showed that the liquid-phase mass transfer coefficient
could be calculated thus:
1/ 3
kl d b  d b3 g ( ρ  ρ g ) 
N Sh,l  2  0.31 N Ra
1/3
or  2  0.31 
D g-l  μ D 
 g-l 
where
Dg-l = gas diffusivity in the liquid phase (m2/s)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Interfacial Area and Gas Holdup


Since the mass transfer coefficient is only a function of fluid properties, then for
any given fluid, the mass transfer rate constant may be correlated simply as a
function of power per unit volume and superficial gas velocity or gas volume
rate, thus:

α α
P β P β
kl a    vs or k l a    Q g
V  V 

where
0.1  α  0.9
0.3  β  0.6

depending on system geometry and fluid properties.


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Typical results of kl a vs P/V


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

EXAMPLE 4 – Estimating the gas-liquid mass transfer rate constant.

Estimate the gas-liquid mass transfer rate constant and the gas holdup for the
pressure oxidation autoclave scenario given in Example 3. Cast the results into a
correlation in specific impeller power and superficial gas velocity assuming a
tank cross-sectional area of 13 m2. Assume a slurry density of 1100 kg/m3 and
ignore the contribution of surface aeration.

Solution. We will use Calderbank’s correlations to estimate the mass transfer


rate constant. However, first we need to estimate a host of fluid properties at
220°C (see course notes for details).

Superficial gas velocity is calculated using the gassing rate from Example 3, thus:

Qg (0.271) m
vs    0.021
S (13) s
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

-0.2

v s = 0.050
:
-0.4 :
Best fit slope = 0.668 0.015
0.010
-0.6 0.005
log kla (s )
-1

-0.8

-1.0

-1.2

-1.4
2.6 2.8 3.0 3.2 3.4 3.6 3.8

log P /V (W/m3)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

-0.2

Best fit slope = 0.273


P /V = 5000
-0.4 :
:
2500
-0.6 2000

1500
log kla (s )
-1

-0.8 1000

-1.0 500

-1.2

-1.4
-2.4 -2.2 -2.0 -1.8 -1.6 -1.4 -1.2
log v s (m/s)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

-1.3

v s = 0.050
-1.5
:
:
Slope = 0.4
0.025
0.020
-1.7
0.015
0.010

-1.9
log

0.005

-2.1

-2.3

-2.5
2.6 2.8 3.0 3.2 3.4 3.6 3.8

log P /V (W/m3)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

-1.3

P /V = 5000
-1.5
:
:
Slope = 0.5
2000
-1.7 1500
1000

500

-1.9
log

-2.1

-2.3

-2.5
-2.4 -2.2 -2.0 -1.8 -1.6 -1.4 -1.2
log v s (m/s)
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

EXAMPLE 4 – (continued)

The correlation which best describes the mass transfer rate constant is:
0.668
P
k l a  0.00356   v s0.273
V 
and the correlation which best describes gas holdup is:
0.4
 P  0.5
φ  0.00488   v s
V 
According to van’t Riet (1979), in non-electrolyte solutions the kl a correlation
exponents are α = 0.4 and β = 0.5, while in electrolyte solutions they are α = 0.7
and β = 0.2. Hence, the value of these exponents seems to depend mostly upon
whether or not the solution contains ions.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Typical results of kl a vs ND/T1/2

CO2-CO32—HCO3- system O2-CuCl system


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Effect of Number and Placement of Impellers

One, two, or three impellers kl a is the same power-law


in various positions and at function of P/V regardless
various depths
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

The Effect of Solids

Correlation of Oguz et al.

kl a decreases with
increasing slurry viscosity
due to the presence of
solids

Probably not an issue in


most hydrometallurgical
applications
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Surface Aeration
Wu (1995) studied surface aeration with a single disk turbine. He found that for
sparging alone in the absence of surface aeration (power in kW):
0.67
P
(k l a ) sp. a.  1.06   v s0.56
V 
while, for surface aeration in the absence of sparging, and with the impeller
placed very close to the liquid surface:
0.65
P
(k l a) su. a.  0.0634  
V 
which leads to the interesting result:

(k l a ) sp. a.
 16.7 v s0.56 > 1 when vs > 0.006 m/s (not that large)
( k l a ) su. a.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

The Mechanism of Surface Aeration


David G. Dixon Mixing and Mass Transfer in Hydrometallurgy
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

The Effect of
Impeller Depth

Results of Wu et al.

kl a increases
dramatically as
impeller breaks
through the surface

Effect less
significant as
sparging rate is
increased
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Gas-Pumping Agitators
Another way to introduce gas into a fluid medium is to pump it in with the
impeller itself. Unlike surface aerators, gas-pumping agitators are well
submerged in the fluid medium.

Gas-pumping can be carried out by two different means:

• Vortexing in non-baffled tanks


• Hollow-shaft impellers

In either case, the key to gas pumping is to overcome the potential energy barrier
represented by gas buoyancy.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Gas-Pumping Agitators
Bernoulli's equation predicts the critical impeller tip speed which signifies the
onset of gas-pumping:

2 gh
ND 
η
where:
h = impeller submersion depth
η = impeller gas-pumping efficiency

Gas-pumping agitators are used extensively in fermentation processes and are


becoming popular in bioleaching applications as well. Gas-liquid mass transfer
rate constants correlate to power per unit volume and the rate of gas-pumping in
the normal fashion.
David G. Dixon Mixing and Mass Transfer in Hydrometallurgy

Application I: Pyrite Autoclaving


The simplest leaching system involving gas-liquid mass transfer is the pressure
oxidation of pyrite, which involves only the following two reactions:

O2(g)  O2(aq) (1-1)

4 FeS2 + 15 O2(aq) + 2 H2O  2 Fe2(SO4)3 + 2 H2SO4 (1-2)

At steady state, the rate of gas-liquid mass transfer must be equal to the rate of
leaching:

r11  k l a C O* 2  C Ob 2  r1 2   4 rFeS2

r1 2 4 rFeS2
rO 2 (aq)  r11  0 Hence C b
O2 C *
O2 
15 15 k l a

The actual rate of pyrite oxidation depends on temperature, particle size, and
dissolved oxygen concentration to the ½-power.

You might also like