Você está na página 1de 6

Ind. Eng. Chem. Process Des. Dev.

1980, 79, 509-514

509

REVIEW
Crystallization Kinetics from MSMPR Crystallizers
John Garside' and Mukund B. Shah
Department of Chemical and Biochemical Engineering, University Co//egeLondon, London WC 1E 7JE, United Kingdom

Published crystallizationkinetics obtained in MSMPR crystallizers are reviewed, and measured nucleation and growth
rates are compared on a common basis. In general, the range of variables studied (supersaturation, magma density,
stirrer speed, etc.) for any given system is extremely limited. Experimental conditions, particularly those related
to crystallizer hydrodynamics, are usually poorly defined and scale-up of kinetics on the basis of such results would
not be possible. Conditions under which future laboratory studies should be made are suggested.

Introduction
The concept of a continuous mixed suspension mixed
product removal (MS MPR) crystallizer (Randolph and
Larson, 1971) has led bo experimental techniques whereby
crystallization kinetics (i.e., nucleation and growth rates)
can be determined under conditions where both of these
kinetic processes are occurring simultaneously. Such kinetics are usually correlated by semiempirical equations
of the form

Table I. Units Used To Compare Kinetic Data

Bo = KRM$G'
(1)
where KR = f (temperature, hydrodynamics, and impurity
concentration) and the relative kinetic order i is the ratio
of the nucleation and growth orders. In the period since
the mid-1960's results of many such experimental studies
have been published for a wide range of systems.
Simultaneously with such experimental studies, development of the population balance approach to crystallizer
design has enabled design equations to be derived for many
crystallizer configurations operating under both steadyand unsteady-state conditions. Before such equations can
be used in a quantitative way the kinetic parameters KR,
j , and i need to be known.
In this paper the results of MSMPR experimental
studies are reviewed and the experimental evidence justifying the use of eq 1 is considered. Data published up
t o mid-1979 have been included. In addition, the possibility of using published kinetic data for crystallizer design
purposes is assessed. The results considered here are restricted to those obtained in MSMPR crystallizers. A
number of published studies contain insufficient detail to
enable quantitative kinetic data to be deduced and results
from these have not been included in this survey.
Results of Kinetic Studies
The literature contains a wide variety of units which
have been employed to record values of the parameters
involved in crystallization studies. In order to compare
published data a common system of units is used
throughout this paper as indicated in Table I.
Table I1 lists the systems that have been studied in
MSMPR crystallizers. The table is arranged according to
the method employed to produce supersaturation and the
ranges of temperature, residence time, and magma density

over which measurements were made are indicated.


Continuous laboratory crystallizers are generally not designed to run at a predetermined level of supersaturation
and so this is not an independent variable under the
control of the investigator. Further, supersaturation levels
are often very low, making them difficult to measure.
Hence only a limited number of studies report a supersaturation range. The growth rate is a measure of the
supersaturation and so the range of G over which kinetics
were determined is also included in Table 11. Published
data were converted to the units shown in Table I and the
resulting correlating equations for Bo are given in Table
11.
In a number of studies the steady-state crystal size
distribution has been measured down to a few micrometers
by using a Coulter Counter (e.g., Youngquist and Randolph, 1972; Randolph and Sikdar, 1976; Garside and
JanEiE, 1979). A rapid increase in population density is
then sometimes observed in these small size ranges and
consequently the derived nucleation rates are very much
higher than when the upward curvature is not present.
Results from such measurements are noted in Table 11.
One further complication arises in comparing kinetics,
that of size-dependent growth. This produces curvature
in the population density plot and a size-dependent growth
rate equation has then to be incorporated into the population balance. Two such equations have been most commonly used, those proposed by Canning and Randolph
(1967)
G = Goc(1 + YCL)
(2)
and by Abegg et al. (1968)

0196-4305/80/1119-0509$01.00/0

variable
nucleation
rate
growth
rate
magma
density

symbo1

unit

Bo

no./L s

m/s

conversion factor

1 no./L s = 0.06
no./cm3 min
lo-* m / s = 0.6 '
pm/min = 36 pm/h

MT g / L = kg/m3

G = GoA(1
0 1980 American Chemical Society

+ YAL)~

(3)

510

Ind. Eng. Chem. Process Des. Dev., Vol. 19, No. 4, 1980

Table 11. Published Crystallization Kinetics from MSMPR Crystallizers

la)
1.

A m m o n i u n alum

Charnblisr 11966)

22

15-45

50-220

3.5-8.3

2.

Ammonium s u l p h a t e

Chambliss (1566)

22

15-45

30- 7 5

4.3-12.2

I.

Ammonium s u l p h a t e

Y a v n g q u i i f and R a n d o l p h ( 1 9 7 2 )

34

5-16

4.

h o n i u n sulphate

Larran and M u l l i n 1 1 9 7 3 )

18

8-20

5.

h o n i v m sulphate

L a r s a n and Klekar ( 1 9 7 3 )

15

9-32

6.

S a g n e i ~ u ms u l p h a t e

S l k d a r and R a n d o l p h (19761

7.

P a r a s s i u n alum

O r f e n s and d e Jong ( 1 9 7 1 )

8. P o r a r E i u m alum

C a r r i d e and J a n E i ;

(1579)

9.

PaLarsium c h l c r i d e

G e n c k and hson
0972)

10.

P o t d l s i ~ mc h l o r i d e

Randolph e t a l .

11.

P o t a s s ~ u nd i c h r o m a t e

T i m and C o o p e r ( 1 9 7 1 )

12.

Potassium dichromare

Desai

11.

Porarsium dichromate

Janse 1 1 9 7 7 )

14.

Polasslum n i t r a t e

S h o r and Larsan (1971)

15.

Potassium n i t r a t e

G e n c k and Larson 119721

16.

P o f a ~ s i u mn i t r a t e

J u i a r n e k and Larsen ( 1 9 7 7 )

17.

Patarslum Illtrace

H e l t and L a r s o n ( 1 9 7 7 )

18.

Potassium su:phafe

R a n d o l p h and Rajugapa:

et

5 -200

IO

10-60

15-120

12-30

15-43

a l . (1974)

(1970)

5.2-11.7

20-100

2.6-10.0

21

15-45

25-87

4 . 8-1 3 . 1

11-40

15-45

25-45

4.3-15.7

20

14-23

10-40

7.7-13.2

..- - I 6

10

19-88

21.

Sodium c h l o r i d e

Bennecr e t a,.

Sodium chloride

Asselbergs (1975)

23,

Citric a c i d

S i k d a r and R a n d o l p h 0 9 : 6 )

24.

Lrea

Bennett and v a n B u r e n ( 1 9 6 9 )

-55

25.

Lrea

Ladaya e t SI.

1-16

,-.

3.7-12.6

3.5-11

5-28

0.3-2.5

0.6-6.1

23-108

1.2-9.5

2.8-7. 1

29

Rose> and H u l b u r c ( 1 5 7 1 )

30

2-10

bO-liC

70-190

0.*8-2.0

50

10-54

25-200

3.5-11

150-4OC

320-110

0.4-1.2

-.

1-7

1.7-1:.8

16-24

1.1-3.7

30-130

Reaction p r e c i p i f a i i o n

26,

Barium n i t r a t e

Blumenfndl e : a l . (1911)

25-45

5-20

27.

Barium soap

Chiiare e t a l . (1576)

15-60

7.5-15

28.

Calcium c a r b a n a r e

S c h i e r h a l z and Stevens ( 1 9 7 5 )

-23

10-1:

29.

Cal'lum

D r a c h e t 8 1 . 119781

10.

Calciua s u l p h a t e !H20

3;.

Calcium sulphate jH20

S i k d a r e l a!.

12.

S i l v e r bromide

Key e t 8 1 .

33.

h o n ~ u alunlH,CiEtOH
r

Yvrray

31.

h o n l u m alun/H,01EfOH

T i m 8nc L e r r a n 11968)

27

35.

Annnar.um

Tim a n d L d r s o n ( 1 9 6 8 )

36.

Soclum rhlor.CeIH,O!ttOH

TIT

37.

S o d l v i c h l ~ - r i d eH 0 l E I O H

Yarns e l 3 1 .

OXdldCe

Salting

38.

1.4-5.2

9-30

R a n d o l p h and S i k d a r 1 1 5 7 6 )

11977)

2-12
1.4-11.0

26-40

10

2-10

11.7-20,i

1.2-3.1

Patassiurr s u l p h a t e

(1971)

0.9-1.5

14-42

10-2s

1,6-I. 3
27.6-67.6

5-1L

18-60

Pacassium sulphate

(11)

14-25
60-250

10-45

19.

'

2.6-7.1
5-50

20.

22.

18-23
9.7-21

18-26

12

(1577)

38
51-61

bmir a n d L a c i o n

(19681

(1978)

11978)

20-120

2.8-18.9

6.00 x I0l6

15-23

5 , 3-11, i

1.72

-0.3

38

5-20

45-75

12-45

0.0:5-0.06
15-18

0 07-0.12

1.6-5.6

70-90

0.20-0.11

80

0.019-0.13

a t 25OC

- ~ , ~ ~

1 . 4 2 x 1021 G 2

0.4-1.1

0.33-1. 7

b$

1 0 - l ~G

2 . 2 8 x 10''
5 . 3 2 x IOz4 G2"
r e a g e n t g r a d e . 7OoC
2.16 x IOz2 G 2 ' 6 p l a n t g r a d e , 70C
8 . 5 9 x 10"

2.2

a c 800C

6 . 3 7 x I O L 9 '6

out

iulphdtelH OIYe0k
2

Sodium i ' h l u r l d a l H O l F f O H

and L a i i o r ( 1 5 6 8 )

64

2.5-8.1

15-45

110

1.4-3.1

15-&5

33

1.7-2.8

I4
11

2 8-9.1

15-:

1.0-5.8

l5-"5

(15721

L i u and B o t s a r i i : I 9 7 3 1

2.9-8.9

15-.5

a n d Larson ( 1 5 6 5 )

27

15-45
c

35.

S o d i u m c h l n r i d c r H OrtlDH

40.

C > c l o l L C i tl\O

,H I:
> 2

5-20

S o n e and D n u p l a s ( . 9 7 5 l

Branrorr e t a l .

11549)

b7

3.1"

2.3-3.9

("I
32.5-53 1

3.5-!4.1

a A dash indicates that the data are not recorded in the original reference.
Size analysis down to 1.26 p m with curvature in population density in this size range; N = 545-675 rpm.
Size analysis down to 8
Effect of Cr3+investigated.
pm, linear population density plot t o this size. N o effect of N o n Bo. e E = 0.1-11.3 W/kg, G o = growth rate of crystals
below 150 pm. f N = 850-2100 rpm. G o= growth rate of crystals below 100 km. Results also for size analysis down t o
5 pm-different kinetics then obtained.
Correlations also given for e = 1 2 and 30 "C.
Solutions near-saturated with
Effect of several impurities investigated.
NaCl and containing 0.75 mg of MgS0,/100 g of H,O.
N = 800-1250 rpm.
Correlation also given for e = 11, 31, and 40 "C.
Size analysis in range
Correlation also given for e = 1 0 and 2 5 "C.
10-30 pm, linear population density plot in this range.
Size analysis down t o 7 pm, nonlinear population density plot,
G o = growth rate in equation of Abegg at al. (1968),(eq 3).
TIPS = impeller tip speed (ft/min), TO = s/turnover, conditions probably not MSMPR. P N = 414-804 rpm. Q N = 600-900 rpm.
For stoichiometric CaSO, and Na,CO,. Results
also for excess Na,CO,.
From phosphoric acid. t Precipitation in gelatin. Size analysis in submicrometer range.
Ci =
concn PbCl, (ppm). Range of Ci = 1-30 ppm.

In both these cases Go represents the growth rate of


"nuclei" and it is this growth rate that is correlated with
the nucleation rate to obtain the relative kinetic order.
Potassium alum and potassium sulfate are systems which
appear to exhibit size-dependent growth.

Figures 1 and 2 depict the crystallization kinetics tabulated in Table I1 as plots of Bo against G. Many of the
studies employing cooling or evaporative crystallization
have investigated the influence of magma density. Where
the effect of this variable has been included in the corre-

Ind. Eng. Chem. Process Des. Dev., Vol. 19, No. 4, 1980 511
,

1os7

105

VI

L
L

a3

G
'm

10'.

-C

Q
0

z' i o 3 -

10'

10-9

,
10-8

!7,,:/4

,
10-7

1u
9

1o

1 o.8
Growth r a t e , G

-~

1 o-6

Irnisi

Figure 2. Crystallization kinetics for precipitation and salting out


crystallization. Kinetics for a given system are represented by lines
of a given form. Numbers indicate reference number in Table 11.
Notes: curve 38, Ci = 1 ppm.

lation the curves in Figure 1 have been drawn for MT =


50 g/L if this value is within the range investigated. Such
a magma density is at the lower end of those generally used
commercially. If the investigation did not extend to this
value, the experimental magma density closest to 50 g/L
was used to calculate a kinetic curve. In cases where the
effect of hydrodynamics has been explored a value of
stirrer speed or power input near the middle of the measured range has been used to evaluate the curve, actual
values being given in the figure key.
Effect of Variables; on Crystallization Kinetics
1. Systems Studied. Choice of systems for use in
academic studies is generally made on the basis of ease of
working and familiarity with the material. In consequence,
many materials of industrial importance are not adequately
represented. For example, no data appear to be available
for ammonium nitrate, the ammonium phosphates, sodium
sulfate, and alumina, while potassium chloride, sodium
chloride, urea, and calcium sulfate have only been the
subjects of a small number of studies. Yet, the annual
U.S.A. production of each of these materials is in excess
of lo6 t (van Damme, 1973). Organic systems are very
poorly represented.
Precipitation processes are being increasingly studied
in MSMPR crystallizers and most of the systems listed in
Table I1 are encountered on the industrial scale. Much
of the early experimental work using the MSMPR technique was carried out using salting-out since this has many
experimental advantages. There appears to be increasing
industrial interest in such processes as a means of increasing yields and conserving energy.
2. S u p e r s a t u r a t i o n , Supersaturation within a
MSMPR crystallizer is forced to different levels by
changing the residence time. As a consequence, the growth
rate changes and this is generally taken as a measure of
the supersaturation as implied by the use of eq 1 to cor-

relate kinetics. The range of supersaturation, and hence


growth rates, that can be achieved by a given change in
residence time depends on the value of i. It can be shown
(Randolph and Larson, 1971) that for constant magma
density
G

a 7-4/(i+3)

(4)

If growth rate is related to supersaturation by the expression


G

0:

sg

(5)

the corresponding change in supersaturation is


s

a ?-4/g(i+3)

(6)

Assuming that the maximum ratio of residence times


that can be achieved in a given crystallizer is 8 (e.g., from
say 7.5 to 60 min) the resulting ratio of G is 8 and 5.3 for
i = 1 and 2, respectively. If g = 1 the corresponding ratio
of s is the same as that for G, while if g = 2 the ratio of
s is 2.8 and 2.3 for i = 1 and 2, respectively. With the
exception of the data for barium nitrate (Blumenthal et
al., 1974, ref 26) for which G varies by a factor of 13.9, the
maximum ratio for G is about 8 but most sets of data cover
a considerably smaller range.
The range of driving force over which kinetics may be
obtained is thus rather limited in practice and this sets a
limit on the accuracy with which kinetic parameters may
be determined.
With few exceptions, values of G measured in published
studies are within the range 2 X
to 2 X lo-' m/s
(- 1-10 pm/min) and are remarkably insensitive to the
particular systems. Precipitation conditions tend to produce growth rates that are somewhat lower.
Nucleation rates cover a much wider range of values. In
part, this must be due to the large number of variables that
affect nucleation rate and that are not standardized in
Figures 1 and 2; magma density and hydrodynamics are
the most important of these. For cooling and evaporative
conditions Bo appears to cover about three orders of

Ind. Eng. Chem. Process Des. Dev., Vol. 19, No. 4, 1980

512

Table 111. Effect of Temperature on Kinetics


ref
(see

Table
111

system
potassium

range
of e, C

i = const., K R

chloride

13
15
17
23

comments

12-30

increases with e

potassium
dichromate
potassium
nitrate
potassium
nitrate

26-40

i = const., B o a

11-40

i = const., K R

10-25

i = const., negative

citric acid

16-24

0 -0.57

decreases with e

activation

energy for B o
i = const., negative

activation

26
30

barium nitrate
calcium

sulfate-water
31

calcium

sulfate-water

energy for B o
25-45
45-75

B o decreases with

70-90

increasing G
i = const., positive

activation

energy for B o

magnitude, from about 2 x 102 to 2 X 105 no./L s ( 10-10~


no./cm3 min). Results of several studies are not included
in Figure 1. Youngquist and Randolph (1972, ref 3) and
Randolph and Sikdar (1976, ref 20) determined steadystate crystal size distributions down to a few micrometers
with a Coulter Counter. Sharp upward curvature was
observed and so the resulting kinetics are not strictly
comparable. Lodaya et al. (1977, ref 25) provide insufficient details of experimental conditions to enable a curve
to be constructed.
For precipitating systems values are much more variable,
covering a least four orders of magnitude. The results of
two studies are not included in Figure 2, Wey et al. (1978,
ref 32) with silver bromide for which nucleation rates are
of the order of 10O to l O I 4 no./L s, and Koros et al. (1972,
ref 37) where nucleation rates appear very low, -lo2 no./L
N

S.

The relative kinetic order for cooling and evaporative


conditions is generally between 1 and 2. The only recorded
values significantly higher than 2 are for KCl where i is
reported as 2.55 (Genck and Larson, 1972, ref 9) and 4.99
(Randolph et al., 1977, ref 10). For potassium dichromate
and citric acid, i appears to be between 0 and 1. Negative
values of i are probably suspect since it is difficult to
envisage kinetic mechanisms giving rise to such values. As
pointed out by Keight (1978), operation of a high yield
MSMPR crystallizer a t constant residence time and
magma density forces the nucleation and growth rates to
follow the relation Bo a G-3. The resulting apparent value
of i = -3 is then unrelated to the kinetics of the system
but arises since too small a range of residence time and
magma density was covered. Nevertheless, there is a
suggestion that potassium sulfate and urea exhibit unusual
kinetic behavior.
With precipitation and salting-out i tends to be higher.
In these systems primary as opposed to secondary nucleation is likely to be the dominating mechanism. Primary nucleation rate expressions are highly nonlinear in
supersaturation and so a high value of i may be expected.
3. Temperature. Few studies have investigated the
effect of temperature on crystallization kinetics. Table I11
summarizes those for which sufficient data are available
to draw tentative conclusions.
The range of temperature over which measurements
have been made is very limited, particularly if attempts
are made to determine activation energies. Apart from
studies with CaS04J/2H20from phosphoric acid, all work

has been conducted close to ambient conditions. Over


these narrow temperature ranges the evidence suggests
that i can be considered constant. In many cases the
apparent activation energy for the nucleation process appears to be negative, and Wey and Terwilliger (1980)
discuss why such results may arise.
4. Magma Density. Secondary nucleation rates which
are removal limited (Evans et al., 1974) will be influenced
by the amount of crystals in suspension. A first-order
dependence on magma density is predicted if crystal-agitator or crystal-crystallizer collisions are responsible for
secondary nuclei production (Ottens and de Jong, 1973).
Most results quoted in Table I1 show such a relation, although it should be noted that in a number of cases the
relation was assumed rather than determined from experimental results. Several correlations show an exponent
on magma density that is close to 0.5 (Desai et al., 1974,
ref 12; Juzaszek and Larson, 1977, ref 16; and two studies
with potassium sulfate, Randolph and Rajugopal, 1970, ref
18, and Randolph and Sikdar, 1976, ref 20) while the study
of Randolph et al. (1977) with potassium chloride gives the
very low exponent 0.14. These low values may arise since
the nucleation is in part regeneration limited (Evans et
al., 1974) or survival limited (Randolph and Sikdar,
1976).
Precipitation studies have not included measurements
on the influence of MT. This reflects the assumption that
primary, rather than secondary, nucleation dominates the
behavior of such processes.
The range of magma density over which measurements
have been made is frequently rather limited and in many
cases is far below typical industrial values. Since crystal-crystal collisions are predicted to result in a secondorder dependence of nucleation rate on magma density,
their relative importance would be expected to increase
at high magma densities. Little experimental work has
been undertaken to test this prediction.
5. Crystallizer Hydrodynamics. Crystallizer hydrodynamics are likely to exert a profound effect on both
primary and secondary nucleation rates. Lack of standardization and poor specification of hydrodynamic conditions are the most likely causes of differences between
kinetics measured by different workers for the same system.
The influence of hydrodynamics on secondary nucleation
was first discussed by Ottens et al. (1972). They showed
that for crystallizers of similar geometry operating in the
removal limited regime Bo 0: 6 which for a given vessel size
implies Bo 0: IP,As shown in Table IV this relation is
approximated in a number of studies. More recent modifications to this theory as summarized by Garside and
Davey (1980) do not give rise to substantially different
conclusions. Several studies show nucleation rates to be
significantly less sensitive to stirrer speed which may again
reflect regeneration or survival limitation, while the very
high order dependence for ammonium sulfate (Youngquist
and Randolph, 1972, ref 3) may be related to the method
of determining the nucleation rate in the micrometer size
range.
The range of stirrer speeds covered in any single investigation is generally very narrow. The minimum stirrer
speed is set by the necessity to suspend all the crystals and
maintain a mixed suspension while the upper limit is set
by vortex formation and air entrainment. This imposes
a limit on the precision with which the exponent of N can
be determined.
Since primary nucleation rates are very sensitive to supersaturation, mixing of the reacting components in pre-

Ind. Eng. Chem. Process Des. Dev., Vol. 19, No. 4, 1980 513

Table IV.

Effect of Civstallizer Hydrodynamics o n Nucleation Kinetics

ref (see
Table 11)
3
6
7
8
13
20
21
22
23
25
31

system
ammonium sulfate
magnesium sulfate
potash alum
potash alum

correlating
relation for B o

potassium dichromate
potassium sulfate
sodium chloride
sodium chloride
citric acid
urea
calcium sulfate

NO

system

4 ammonium
sulfate
1 4 potassium
nitrate

29

calcium
oxalate
38 sodium chloridewater-ethanol
- borax (Randolph
and Koontz,
1976)
- potash alum
(Rousseau and
Woo, 1978)

N 3 ;results for 1 0 and 40 L


for B o determined with Coulter Counter,
B o a N";results for 1, 5, and 3 0 L

E a
h1.8

N3
N2.5

TIPS2/T0
N2

TIPSZ/TO a N 3
55-L vessel

NO
1V 2.3

NO

Table V. Effect of Impurities o n Kinetics


ref
(see
Table
111

comments

N7.84

impurity and
concentration range
Cr3+,0-20 ppm
Co2+,0-4 g/100 mL soln
Cr3+,0-500 ppm
methylamine hydrochloride,
0-1000 ppm
dodecylamine hydrochloride,
0-50 ppm
fluorocarbon, 0-30 ppm
pyrophosphate
methylene blue
Pb", 1-30 ppm
sodium oleate
dodecylbenzene sulfonate,
0-18 ppm
quinoline yellow,
0-300 ppm

cipitation and salting out systems may be of importance.


Local regions of high supersaturation near the mixing point
could produce extremely high nucleation rates which would
be very sensitive to thLe degree of mixing and hence stirrer
speed. None of the studies reported in sections (ii) and
(iii) of Table I1 adequately investigated these problems.
6. Impurities. The dramatic effect of trace quantities
of impurities on crystallization kinetics is well known.
Little attention has been paid, however, to quantitative
investigations of their influence in MSMPR crystallizers.
In large part, this is 11-10doubt due to the lack of an adequate theoretical framework on which to base experiment
planning and interpretation. Table V lists those studies
that have been reported and includes two references where
insufficient data are available to determine kinetics of the
pure system and so are not included in Table I1 (Randolph
and Koontz, 1976; Rousseau and Woo, 1978). Increasing
impurity concentration generally decreases both nucleation
and growth rates for a given supersaturation. In an
MSMPR crystallizer EL decrease in, say, nucleation rate will
result in a decrease iin surface area and so a rise in supersaturation. The imam balance constraint, therefore,
forces the nucleation and growth rates to move in opposite
directions, and the separate responses of nucleation and
growth kinetics are difficult to disentangle.
The evidence available at present suggests that i is
relatively independent of impurity concentration although
KR can be very sensitive to small changes in impurities.
Resulting crystal size distributions can hence be changed

significantly by impurity addition.


Discussion and Conclusions
Although it was not the objective of many of the studies
reported here to obtain crystallization kinetics that could
be related to crystallizer design and operation, such an aim
is important if the power of the population balance to
explain crystallizer behavior is to be exploited. The relative
kinetic order, i, is probably the single most important
parameter and for several systems the values determined
by different workers are encouragingly similar: e.g., ammonium sulfate, i = 1.03 to 1.5; potash alum, i = 1.58 to
1.8; potassium dichromate, i = 0.48 to 0.53, and CaS04l/zHzO,i = 2.8 to 3.2. For some other systems there are
large differences: e.g., potassium chloride, i = 2.55 to 4.99;
potassium sulfate, i = -1 to 0.54 and sodium chloride/
H,O/EtOH, i = 1.72 to 9.0. The absolute value of the
nucleation rate for a given growth rate frequently varies
by several orders of magnitude as with, for example, potassium dichromate, potassium nitrate (see Figure l ) ,
CaS04J/2H20and some of the salting out systems (see
Figure 2). Some possible reasons for this have already been
discussed. It should also be noted that the volumetric basis
on which nucleation rates are reported is not always clear
in the original references.
The basic form of the kinetic equation (eq 1) seems
adequate to represent experimental data but the ranges
of magma density and hydrodynamics over which a given
study have been made are in general very limited. The
frequently accepted value of j = 1 needs more extensive
experimental support, particularly at high magma densities
and for large-scale crystallizers, and the effect of mixing
in precipitating and salting-out systems needs further investigation. Hydrodynamics of crystallizers used for kinetic studies need to be more clearly defined by the investigators. A useful step would be to record power numbers for the agitator/crystallizer system used in kinetic
investigations since these could then be related directly
to the specific power input as suggested by the nucleation
model of Ottens et al. (1972). Standardization of vessel
and stirrer geometry would be of the greatest value. It is
probably true to say that few of the data recorded in Table
I1 could be used with any conficence for crystallizer design
purposes. Finally, comparison between different sets of
data would be facilitated if a common system of units were
used for the important crystallization parameters.
Nomenclature
b = exponent in size-dependent growth equation (eq 3)

Bo = nucleation rate, no./(L s)

g = growth order (eq 5)


G = growth rate, m / s
GOA= nuclei growth rate (eq 3) (m/s)

514

Ind. Eng. Chern. Process Des. Dev.

Goc = nuclei growth rate (eq 2) (m/s)

1980, 19, 514-521

Crystallization from Solution", presented at 66th Annual AIChE Meeting,


Philadelphia, 1973.
Larson, M. A., Mullin, J. W., J. Cryst. Growth, 20, 183 (1973).
Liu, Y-A., Botsaris, G. D., AIChE J., 19, 510 (1973).
Lodaya, K. D., Lahti, L. E., Jones, M. L., Ind. Eng. Chem. Process Des. Dev.,
16, 294 (1977).
Murray, D. C., Larson, M. A,, AIChE J., 11, 728 (1965).
Ottens, E. P. K., Janse, A. H., de Jong, E. J., J. Cryst. Growth, 13/14, 500
(1972).
Ottens, E. P. K., de Jong, E. J., Ind. Eng. Chem. Fundam., 12, 179 (1973).
Randolph, A. D., Rajagopal, K., Ind. Eng. Chem. Fundam., 9, 165 (1970).
Randolph, A. D., Larson, M. A., "Theory of Particulate Processes", Academic
Press, New York, 1971.
Randolph, A. D., Sikdar. S . K., Ind. Eng. Chem. Fundam., 15, 64 (1976).
Randolph, A. D., Koontz, S.,"Effects of Habit and Nucleation Modifiers In
Crystallization of Sodium Tetraborate Decahydrate", presented at 69th
Annual AIChE Meeting, Chicago, 1976.
Randolph, A. D., Beckman, J. R., Kraljevich, Z. I.,
AIChE J., 23, 500 (1977).
Rosen, H. N., Hulburt, H. M., Chem. Eng. frog. Symp. Ser. No. 110, 67, 18
(1971).
Rousseau, R. W., Woo, R., "Effects of Operating Variables on Potassium Alum
Crystal Size Distrlbutions", presented at 84th National AIChE Meeting, Atlanta, 1978.
Schierhoiz, P. M., Stevens, J. D., AIChE Symp. Ser. No. 151, 71, 248 (1975).
Sikdar, S.K., Randolph, A. D., AIChE J., 22, 110 (1976).
Sikdar, S. K., Ore, F., Moore, J. H., "Crystailizatlon of Calcium Sulfate Hemihydrate in Reagent Grade Phosphoric Acid", presented at 84th National
AIChE Meeting, Atlanta, 1978.
Shor. S. M., Larson, M. A,, Chem. Eng. frog. Symp. Ser. No. 110, 67, 32
(1971).
Song, Y. H., Douglas, J. M., AIChE J., 21, 924 (1975).
Timm, D. C., Larson, M. A., AIChE J., 14, 452 (1968).
Timm, D. C., Cooper, T. R., AIChE J., 17, 285 (1971).
van Damme-van Weeie, M. A., in "Crystal Growth: an Introduction", Hartman,
P., Ed., North Holland, p 248, 1973.
Wey, J. S., Terwilliger, J. P., Glngello, A. D., "Analysis of Silver Bromide
Precipitation in a Continuous Suspension Crystallizer", presented at 84th
National AIChE Meeting, Atlanta, 1978.
Wey, J. S., Terwllliger, J. P., Chem. Eng. Commun., 4, 297 (1980).
Youngquist, G. R., Randolph, A. D., AIChE J., 18, 421 (1972).

i = relative kinetic order

j = exponent of magma density (eq 1)


KR = nucleation rate constant, no./L s (g/L)I (m/s)'
MT = magma density, g/L
N = agitator speed, rpm
s = supersaturation, g of crystallizing substance/g of water
YA, yc = constants in size-dependent growth equations (eq
3 and 2), rn-l
E = specific power input, W/kg
T = residence time, s
0 = temperature, "C

Literature Cited
Abegg, C. F., Stevens, J. D., Larson, M. A., AIChE J., 14, 118 (1968).
Amin, A. B., Larson, M. A., Ind. Eng. Chem. Process Des. Dev., 7 , 133 (1968).
Asselberg, A. J., Ph.D. Thesis, Delfi Technical University, 1978.
Bennett, R. C., van Buren, M., Chem. Eng. f r o g . Symp. Ser. No. 95, 65, 44
(1969).
Bennett, R. A., Fiedeiman, H., Randolph, A. D., Chem. Eng. frog., 69(No. 7),
86 (1973).
Biumenthai, E., Keight, D. V., Rolfe, N., "Crystallization: Laboratory Tests",
lecture to Institution of Chemical Engineers, Manchester, U.K. Jan 8, 1974.
Branson, S.H., Dunning, W. J., Millard, B., Discuss. Faraday Soc., 5 , 83 (1949).
Canning, T. F., Randolph, A. D., AIChE J., 13, 5 (1967).
Chambliss, C. W., Ph.D. Thesis, Iowa State University, 1966.
Chlvate, M. R., Vaidya, A. M., Tavare, N. S., Indian J. Techno/., 14, 569 (1976).
Desai, R. M., Rachow, J. W., Timm, D. C., AIChE J., 20, 43 (1974).
Drach, G. W., Randolph, A. D., Miiier, J. D., J. Urol., 119, 99 (1978).
Evans, T. W., Margolis, G., Sarofim, A. F., AIChE J.. 20, 950 (1974).
Garside, J., JanEie, S. J., AIChE J., 25, 948 (1979).
Garside, J., Davey, R. J., Chem. Eng. Commun., 4, 393 (1980).
Genck, W. J., Larson, M. A., AIChE Symp. Ser. No. 121, 66, 57 (1972).
Hen, J. E., Larson, M. A., AIChE J., 23, 822 (1977).
Janse, A. H., Ph.D. Thesis, Delfi Technical University, 1977.
Juzaszek, P., Larson, M. A,, AIChE J., 23, 460 (1977).
Keight, D. V., Ind. Eng. Chem. Process Des. Dev., 17, 576 (1978).
Kwos, W. J., Daltymple, D. A., Kuhlman, R. P., Brockmeiet, N. F., AIChE Symp.
Ser. No. 121, 68, 67 (1972).
Larson, M. A,, Kieker, S. A,, "In-situ Measurement of Supersaturation in

Received f o r review January 28, 1980


Accepted April 30, 1980

ARTICLES

Electrochemical Study of Liquid-Solid Mass Transfer in Packed


Beds with Upward Cocurrent Gas-Liquid Flow
Ghlslalne Delaunay, Alain Storck, Andre Laurent, and Jean-Claude Charpentler'
Laboratoire des Sciences du G6nie Chimique, CNRS-ENSIC, 54042 Nancy, France

An electrochemical technique was used for global measurement of overall solid-liquid mass transfer coefficients
in upward cocurrent gas-liquid flow through a packed bed under bubble flow and surging flow conditipns. The
feasibility of the technique was ascertained and the coefficients were compared with those obtained by other
techniques. An energetic correlation has been proposed for both single and gas-liquid flow which extends the
range of application of previous works.

Introduction
The use of fixed bed reactors operated under cocurrent
upflow conditions has widely increased during these past
years, especially in the petrochemical industries: coal liquefaction, catalytic hydrodesulfurization, selective hydrogenations, etc. (Shah, 1979). The overall rate of the
process may depend either on the chemical reaction kinetics or on the physical gas-liquid and liquid to solid
0196-4305/80/1119-0514$01.00/0

particle mass transfer resistances. However, under certain


operating conditions the limit of the process is only located
at the diffusional mass transfer resistance near the liquid-solid interface.
There are very few data available on the determination
of the space_average value of the particle mass transfer
coefficient k for upward cocurrent flow in packed beds.
The recent works published on the topic together with the
0

1980 American Chemical Society

Você também pode gostar