You are on page 1of 6

Light Metals 2009 Edited by: Geoff Bearne

TMS (The Minerals, Metals & Materials Society), 2009

ALUMINA DISSOLUTION AND CURRENT EFFICIENCY IN HALL-HEROULT CELLS


Bjrn Lillebuen1, Marvin Bugge1 and Helge Hie2
Hydro Aluminium, P.O. Box 2560, NO-3908, Porsgrunn, Norway
2
Hydro Aluminium Karmy, NO-4265, Hvik, Norway

Keywords: Alumina, Current Efficiency (CE), Sodium, Dissolution

Abstract
In this paper we will present some operational data, and discuss
correlations and links to the phenomena introduced above.

The dissolution and distribution of alumina in cryolite melts can


be described as a coupled heat and mass transport process, with
intermediate formation of solid cryolite. Current efficiency (CE)
can be evaluated by means of the rate equations for the back
reaction between dissolved metal and carbon dioxide gas. Solid
cryolite may be formed close to the metal pad under certain
conditions in the cell. The bath superheat and the mass transfer
coefficient at the bath/metal interface will be important
parameters for the maximization of current efficiency. In some
cells there is a clear correlation between current efficiency and the
sodium content in the metal, indicating that mass transfer is the
dominant factor. In other cells sodium levels can be quite low
even at high current efficiency, which indicates that formation of
solid cryolite can play a significant role, making superheat the
dominating factor.

Alumina Dissolution
1. Alumina Quality
Alumina quality is often considered to be the main parameter for
alumina dissolution control. Modern sandy alumina has quality
variations that represent challenges for point feeding control, but
more and more we have come to see that these variations may be
less critical than the heat and mass transport conditions inside the
cells. Successful implementation of point feeding has enabled
computer control of the alumina concentration in the bath, and
made possible a dramatic reduction in anode effect frequency and
duration, in addition to the general performance improvement
seen in new smelters. In order to improve from todays situation, a
detailed understanding and description of alumina dissolution is
required.

Introduction
1. Alumina Dissolution

2. Dissolution of an Alumina Particle


When alumina is added to molten bath, it starts to dissolve. The
dissolution process is endothermic, so heat needs to be transported
to the dissolution interface by a driving temperature gradient.
Since the bath temperature normally is just a few degrees above
the bath liquidus temperature, the alumina dissolution may lead to
local formation of solid cryolite. This solid cryolite will then
gradually melt as more heat is tranported to the interface.

For illustration, a simplistic calculation of the dissolution rate of


an alumina particle is shown. The particle is heated to bath
temperature and dissolved, and Figure 1 gives an illustration of
temperature and concentration gradients in the bath boundary
layer (film) around the particle, and related to the bath phase
diagram.

2. Aluminum Electrolysis
The electrolytic deposition of molten aluminum takes place with
sodium ions as charge carrier, and discharging of Al-containing
species at the metal/bath interface. This sets up a concentration
gradient with higher bath ratios at the interface than in the bulk of
the bath. Higher bath ratio means higher liquidus temperature, so
formation of cryolite may take place also at the cathodic interface
between metal and bath.
3. Current Efficiency and Sodium
The sodium levels in aluminum are determined by the bath ratio at
the cathodic interface according to the equilibrium reaction:
Figure 1. Particle Dissolution.
Al + 3 NaF = AlF3 + 3 Na

(1)
The following equations can be written for the dissolution rate,
ref. Asbjrnsen and Andersen (1):

From this same interface metal will dissolve and react with carbon
dioxide by the so-called back reaction. The back reaction is
assumed to be responsible for the largest part of the current
efficiency losses in modern cells. In addition to Al, metals like Na
and Li will also dissolve and back react, and dissolved metals may
also react with impurities like phosphorous in the melt.

389

M = k (c* - c)

(2)

Q = h (T T*)

(3)

Q = M (Hdiss)

(4)

(T T*)/(c*- cs) = const

(5)

From the figure it seems to last about 100 seconds for the
temperature to climb back up to the value shown before feeding

where M is dissolution rate, Q is heat transport rate, Hdiss is total


dissolution enthalpy, and the concentrations can be interpreted by
means of Figure 1. Solving these equations:

M = k (a / (a+1)) (cs c)

(6)

a = (h/k) (const/Hdiss)

(7)

where:

Assuming that a is much larger than 1, then:


M = k (cs c)

(8)

This means the normal rate equation as the product between the
mass transfer coefficient k and the concentration gradient for the
constant bath temperature case.

Figure 2. Alumina dissolution rate, from Thonstad et al . (2).

The differential equation for dissolution of the particle can be


written:
dt = const (Adr)/ (MA)

(9)

where A is area and r is radius. The mass transfer coefficient k can


be estimated from the relation:
Sh = 0.332 (Re)

1/2

(Sc)

1/3

(10)

The dimensionless groups are estimated by inserting relevant


physical data, and assuming an interfacial velocity of 5 cm/s.
Integrating from t = 0 to t and r = r to 0 gives:
t = 26573 (r)

3/2

(11)
Figure 3. Temperature drops during dissolution of 1.5 kg alumina
doses, taken from Kobbeltvedt et al. (3).

For a normal alumina particle with diameter 100 micron, r = 0.005


cm, then:

.
t = 9.4 seconds

(12)
In order to illustrate the thermal effects during dissolution, the
approximate volume of bath needed to dissolve 1.5 kg alumina is
calculated, assuming perfect mixing in the bath, no exchange with
the surroundings, and with ten degrees of bath cooling. With
reference to the works done by Bratland and Grjotheim (4) and
Holm (5), the bath volume (x) can be written as a function of the
temperature drop (dT):

Thonstad et al. (2) have made careful laboratory experiments and


measured dissolution rates in the range of 5-10 seconds, which is
in good agreement with our very approximate calculation for a
constant temperature dissolution. Some of Thonstads data are
reproduced in Figure 2.

3. Dissolution under Point Feeders


Kobbeltvedt et al. (3) have shown that bath temperatures
measured near the feeder will drop several degrees every time the
feeder adds 1.5 kg alumina, see Figure 3, taken from his work.

dTx = 494

(13)

x = 49 kg bath

(14)

or:

390

where r is back reaction rate = metal dissolution rate, k is mass


transfer coefficient for dissolved metal, A is metal/bath interface
area, cAl is metal saturation concentration at the interface
metal/bath, and f1 and f2 are functions of mass transfer parameters
and solubility parameters for the metal/bath and for the gas/bath
interfaces.

In words: 1.5 kg alumina will dissolve with a cooling effect


corresponding to ten degrees cooling of 49 kg bath.
Two comments should be made. Firstly, the calculation above was
made for a bath with 2.3 wt% alumina. The dissolution enthalpy is
strongly dependent upon alumina concentration as shown in
Figure 4, taken from Holms early work (5). Cryolite formation is
therefore favored if the cell operates at low alumina
concentrations in the bath.

If the gas bubble area (Ab) is very small, the following


approximation can be written:

rAl = kAl Ab cAl

Secondly, it is conceivable that some of the alumina/cryolite mix


can float on the metal surface, thus influencing on-going
processes at the interface. This will be discussed in the following
with a view on current efficiency and the sodium content in Al.

(16)

If instead the bubble area is very much larger than the metal area,
then:

rAl = kAl AAl cAl

(17)

For practical application in plant cells we may write finally:

rAl = kAl AAl cAl (1 f )

(18)

Here f can be treated as a cell calibration constant, and:


CE = 100% (r rAl) /r

(19)

when r is metal production for CE = 100%.


It is possible to describe also the effect of bath impurities on
current efficiency in a similar context. If the bath contains an
n+
oxidized impurity i which reacts with dissolved metal in the
bath boundary layer close to the metal surface, then:
i

n+

+ (n/ 3) Al = i + (n/ 3) Al

3+

(20)

This will enhance the metal dissolution rate, thereby reducing


current efficiency, with an enhancement factor (e), given by:
e = 1 + (n/ 3) (Di / DAl) (ci/cAl )
Here D is diffusivity for i
n+

n+

(21)

and for dissolved Al, respectively, ci

is concentration of i in the bulk of the bath, and cAl is metal


solubility in the bath at the metal/bath interface. Equation (21) can
help explain why some multivalent impurities like phosporous can
reduce the current efficiency.

Figure 4. Partial Dissolution Enthalpy, from Holm (5).


.

Sodium and Phosphorous in Al


Current Efficiency

1. Sodium in Al

Current efficiency is reduced in a number of ways, for instance by


electronic conduction, and by the presence of impurities like
phosphorous.

The well-known shift in bath composition passing from bulk bath


through the boundary layer and on to the metal/bath interface, can
be described by (6):

However, it is expected that current efficiency reductions mainly


occur as a result of the back reaction between dissolved metal and
carbon dioxide gas. Applying boundary layer diffusion theory
(classical film theory) it is easy to show that:

rAl = kAl AAl cAl (1 f1/f2)

CAlF3 (bulk) CAlF3 (interface) = I / nFk

(22)

where F is Faraday constant, I is the current density, k is the mass


transfer coefficient, and C is the concentration.

(15)

The high bath ratio at the interface will determine sodium levels
in Al according to:

391

Older pots show lower sodium content, displayed in Figure 7,


which is in line with the normal reduction of current efficiency
with increasing age, shown in Figure 9.

(23)

Al + 3NaF = 3Na + AlF3

In this way, sodium in Al will depend on mass transfer and


current density, and less on bulk bath ratio, see Tabereaux (7).
In Figure 5 the sodium in Al is plotted along with current
efficiency for one side-by-side potline, as time series over several
years. In spite of many different operational upsets, the
covariation between sodium and current efficiency is obvious.
Sodium will sometimes change before current efficiency, as
expected since current efficiency numbers are based on aluminum
tapping weights.
No similar covariation has been found for older end-to-end and
Soederberg lines. Sodium levels are normally lower in these older
lines, even though current efficiency may sometimes be higher.

97
CE

100

95

80

93
Na

60
40

CE %

Na ppm

120

91

Excellent current efficiency values can be achieved with low


noise levels, as indicated in Figure 8.

89

Month

Figure 7. Sodium in Al as a function of cell age (days).

Figure 5. Current efficiency and sodium content, monthly values.

Results of a comprehensive study in one of our potlines are


displayed in Figures 6-8. All cell numbers were averaged over one
whole year, and a regression analysis was performed:

Na (ppm) = -1271 + 14.4 (%CE)

(24)

This explained 84% of the data variation displayed in Figure 6.


.

Figure 8. Current efficiency and cell noise.

In Figure 9 current efficiency is plotted versus cell age. These data


have been collected some years later than data in Figures 6-8, but
confirm the trend.

Figure 6. Sodium in Al versus current efficiency, yearly cell data.

392

98

indicates that well-balanced magnetic fields (low k) as well as


high current density (low A) should both increase current
efficiency. As is well known, this is not always happening, and
some of the complicating factors will be discussed in the
following.

Current Efficiency (%)

97
96
95
94

Reduced turbulence will reduce k, but it will also increase c, since


metal solubility will increase when the bath ratio increases, see
Peterson and Wang (8). We have tried to quantify the relation
between k and c, and our preliminary conclusion is that changes in
k will be modified by changes in c, but the turbulence factor will
always remain dominant.

93
92
91
90
89
0

500

1000

1500

2000

2500

3000

3500

Pot Age (days)

Figure 9. Current Efficiency (one year average) and cell age.

Electronic conduction is known (9) to increase in importance as


bath ratio increases, but it is not clear if the interface ratio or the
bulk bath ratio will determine the effect of electronic conduction
on current efficiency
Higher bath ratio at the interface will increase the possibility for
cryolite to precipitate, thereby reducing the metal area and
reducing back reaction rate. Obviously there will be a limit above
which excessive cryolite formation can lead to erratic current
distribution and even to bottom sludge.

2. Phosphorous in Al
A detailed study for phosphorous, partly displayed in Figure 10,
gave the following regression correlation:

Cryolite can form directly at the metal interface, or during


alumina dissolution. In both cases bath superheat will be an
important parameter. Dewing (10) published the following
empiric regression equation:

P (ppm) = 19.4 0.0052 (A - 1130) + 1.6 (CE 93.8) (25)


A is cell age in days. This equation indicates that the phosphorous
level in Al will increase with increasing current efficiency, but
equation (25) could only explain 40% of the scatter.

Log (%CE) = 0.0095 dT 0.019%AlF3


0.06%LiF + const.

(27)

Lower superheat (dT) will according to this equation lead to


higher current efficiency, and cryolite formation may be a
possible part of the explanation.
High current density will increase heat production in the cell. In
order to conserve energy, every pcell designer and cell operator
will strive to reduce the anode-cathode distance (ACD). This will
eventually increase turbulence, with deteriorating current
distribution and increased gas release resistance. The reduced
ACD may also result in more anode problems like anode spikes,
and reduce the current efficiency. To some extent the low ACD is
today compensated by multiple and deep slots in the anodes.
The high bath ratio at the metal interface will also tend to increase
the Marangoni turbulence, which has been described by Utigard
(11). Interfacial tension between bath and metal varies with the
bath ratio, and this creates horizontal velocities in the bath
boundary layer due to bath ratio gradients. The discussion of these
phenomena also involves the slow transport of alumina (and
fluoride) from the cell bottom to the electrode space, through the
bath film assumed to surround the entire metal pool in the cell.

Fig. 10. Phosphorous in Al versus cell age.

A1 and A2 represent two different anode qualities with slightly


different phosphrous content.

The idea that cryolite is an active ingredient in the cell processes


is not a new one. Recently, Solheim (12) modelled the conditions
for cryolite formation, and Haupin (13, 14) has mentioned cryolite
precipitation as a possible explanation for the so-called liqidus
enigma found in many cells. The measured resistance hysteresis
reported by Kvande and co-workers (15) can perhaps also be
explained by means of cryolite precipitation.

Discussion
The standard equation for back reaction rate:

rAl = kAl AAl cAl

(26)

393

In order to maintain excellent current efficiency and energy


consumption at very high electrode current density in large cells,
it may be necessary to optimize both alumina feeding (16) and the
ACD in a more localized way than before, in order to control
cryolite formation.

5.

6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.

Conclusion
Alumina dissolution in cells with point feeders proceeds with
intermediate cryolite formation.
To melt, dissolve and distribute the alumina-cryolite mixture,
sufficient turbulence (locally) is needed.
High current density, low superheat and good magnetic stability
may lead to precipitation of cryolite on the metal surface.
Some amount of solid cryolite formation can increase the current
efficiency by slowing down the back reaction rate, but excessive
amounts of cryolite could disturb current distribution, form
bottom sludge and reduce current efficiency.
Individual control of point feeders and anodes may benefit the
development of large cells with high anodic current density, as
discussed by Moxnes (16).

Picture 1. HAL275 in Sunndal, Norway.

References
1.
2.
3.
4.

O.A.Asbjrnsen and J.A.Andersen, Light Metals 1977,


pp. 137-143
J.Thonstad et al., Met. Trans. 3, p. 403, (1972)
O.Kobbeltvedt et al., Light Metals 1996, pp. 421-428
D.Bratland and K.Grjotheim, Light Metals 1976, Vol. 1,
pp. 3-21

394

J.L.Holm, Thermodynamic Properties of Molten


Cryolite and other Fluoride Mixtures, Inst. of Inorganic
Chemistry, NTNU, Trondheim, Norway, 1971
J.Thonstad and S.Rolseth, Electrochimica Acta, 1978,
Vol. 23, pp. 233-241
A.T.Tabereaux, Light Metals 1996, pp. 319-326
R.D.Peterson and X.Wang, Light Metals 1991, pp. 331337
G.M.Haarberg et al., Light Metals 2002, pp. 1083-1089
E.W.Dewing, Met. Trans. B, Vol. 22, 1991, pp. 177-182
T.Utigard et al., Light Metals 1991, pp. 273-281
A.Solheim, Light Metals 2002, pp. 225-230
W.Haupin, Light Metals 1992, pp. 477-480
W.Haupin, Light Metals 1997, pp. 319-324
H.Kvande et al., Light Metals 1997, pp. 403-410
B.P.Moxnes, Light Metals 2009, in print.

You might also like