You are on page 1of 5

Ind. Eng. Chem. Fundam.

1984, 23, 143-147 143

Z =
composition in W phase, mole fraction Literature Cited
Cassamatta, G.; Bouchey, D.; Angelina, H. Chem. Eng. Sci. 1978, 33, 145.
Subscripts and Superscripts King, C. J. "Separation Processes", 2nd ed.; McGraw-Hill: New York, 1980.
I, II, III = phase boundary Li, N. N. U.S. Patent 3 650 091, 1972.
arb = arbitrary Misic, D. M.; Smith, J. M. Ind. Eng. Chem. Fundam. 1971, 10, 380.
Wankat, P. C. Ind. Eng. Chem. Fundam. 1980, 19, 358.
E = equilibrium Ward, W. J., Ill AIChEJ. 1970, 16, 405.
in = inlet Way, J. D.; Noble, R. D.; Flynn, T. A.; Sloan, E. D. J. Membr. Sci. 1982, 12,
out = outlet 239.

Greek Letters Received for review November 29, 1982


p =
density, mol/m3 Accepted August 15, 1983

Kinetics of Phase Transitions in the System Sodium Sulfate-Water


David Rosenblatt,Stephen B. Marks,$ and Robert L. Plgford**
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Institute of Energy Conversion and Department of Chemical Engineering, University of Delaware, Newark, Delaware 19716

The rates of growth and solution of single crystals of sodium sulfate decahydrate and anhydrous sodium sulfate
Downloaded via GAZI UNIV on February 12, 2024 at 07:03:53 (UTC).

were measured under conditions of low driving force and low solution velocity past the crystal. The rates of growth
and solution of sodium sulfate decahydrate were found to be controlled by the rates of heat and mass transport
between the crystal surface and the bulk solution. The rate of growth of anhydrous sodium sulfate was controlled
by the rate of incorporation of ions into the crystal lattice at the crystal surface. This information can be used
to predict the rates of phase transitions in a system containing sodium sulfate and water, where the crystal
populations, nucleation rates, and the fluid mechanics at the crystal surfaces are known.

Introduction kinetics has been undertaken in order to determine which


Sodium sulfate decahydrate (Glauber’s salt) is an at- factors control the rates of solution and crystallization of
tractive material for use as a phase change thermal energy Glauber’s salt.
storage material. Its melting point (32.3 °C) and enthalpy The Sodium Sulfate-Water System
of fusion (60 cal/g) make it potentially capable of storing Figure 1 shows a partial phase diagram of the system
solar thermal energy for residential heating purposes over the range of temperatures and concentrations of in-
particularly in so-called “passive” installations. Thermal terest. A detailed description of the Na2S04-H20 system
energy is stored by using solar energy to melt the phase- can be found in the work of Wetmore and LeRoy (1951).
change material; later, when the solar energy is no longer Curve AB represents the solubility of Na2SO4-10H2O as
available, the melted material crystallizes and releases the a function of temperature, BC is the solubility of Na2S04
stored energy. In order to be of use in a practical system, vs. temperature, and DE is the composition of pure Na2-
the phase-change material must possess certain properties: SO4-10H2O. The intersection of AB and BC represents the
suitable phase-change temperature, large heat of fusion, “melting point” of the decahydrate; however, since B is to
and sufficiently rapid rates of melting and crystallization. the left of D, the decahydrate is said to “melt” incon-
The first two criteria are somewhat obvious; however, gruently; i.e., a crystal of decahydrate decomposes to give
the importance of the third point has not always been solution and excess Na2S04 in the ratio 10 mol of H20 to
appreciated. Many solar thermal energy storage systems 1 mol of Na2S04. A kinetic study of the system is com-
must be capable of storing and releasing energy in a 24-h plicated by incongruent melting of the decahydrate at
period; therefore, the phase change material must be able compositions to the left of E. In order to simplify the
to melt and crystallize within this time. Generally, the rate project, it was decided initially to study the rates of
of melting is a function of the heat and mass transfer crystallization and solution of individual crystals of Na2S04
characteristics of the material; however, for many materials and Na2SO4-10H2O under varying conditions of tempera-
the rate of crystallization is often controlled by a chemical ture, supersaturation, and solution flow rate past the
reaction at the crystal surface. Thus, the rate of crys- crystal surface.
tallization may be limited by factors other than mass and
heat transfer. In such a case, the enthalpy of crystalliza- Experimental Section
An apparatus was constructed in which a sodium sulfate
tion, i.e., the stored energy, cannot be released faster than
the material can crystallize and so this process becomes solution, whose concentration and temperature were pre-
rate determining. It is critical to determine the factors cisely adjustable, flowed at a known velocity over a single
governing the crystallization and melting kinetics of any crystal of either Na2S04 or Na2SO4-10H2O. The solution
was preconditioned in a large holding tank so that the
candidate phase change material such as Glauber’s salt.
desired degree of supersaturation was obtained. The
Therefore, a systematic study of the system’s phase-change
conditioned solution was pumped at a known rate over the
crystal which was suspended from a glass rod in a glass

Institute of Energy Conversion. tube. A thermocouple was suspended in the tube near the
5
Department of Chemical Engineering. surface of the crystal. The solution was then pumped into
f E. I. du Pont de Nemours and Co., Inc., another holding tank in which the temperature was de-
Richmond, VA.

0196-4313/84/1023-0143$01.50/0 © 1984 American Chemical Society


144 Ind. Eng. Chem. Fundam., Vol. 23, No. 2, 1984

Figure 1. Partial phase diagram of the sodium sulfate-water sys-


tem.

-0.30 0.20 -0.10 0 0.20 0.30


liberately set to undersaturate the solution so that incipient 0.10

nuclei or crystallites were prevented from being circulated (Cb-C.) (Cb-C,l


UNDERSATURATION SUPERSATURATION
through the system. The solution then was filtered and ( g mola/liter ) ( g mole/liter )

returned to the first holding tank for reconditioning and 2. Variation of crystal growth and solution rate with
Figure con-
recirculation. At periodic intervals the suspended crystal centration driving force.
was removed from the flow tube and weighed. In this
manner data concerning the rates of growth and solution The rate-limiting step in the process was the rate of heat
of Na2S04 and Na2SO4-10H2O were generated. and mass transport between the crystal and the bulk so-
It is estimated that, because of the large volume of so- lution.
lution (ca. 35 L), temperature fluctuations were less than The flux of sodium sulfate to the interface is given by
0.05 °C at flow rates of up to 1.0 L/min. The solution
temperature was measured within 0.1 °C by an iron-
constantan thermocouple located near the suspended Ns =
kCtF In (4)

crystal. Solution concentrations were determined by


pycnometry within ±0.005 M. All solutions were prepared In this expression, derived from diffusion film theory, F
from Baker’s analyzed reagent grade anhydrous Na2S04 NJ (Ns + Nw) is the ratio of the flux of salt to the total
=

and distilled water. The velocity of the solution was de- flux of material and X; and Xb are the interface and bulk
termined by a rotameter placed in series with the crystal mole fractions of Na2S04. This mass transport expression
cell. takes into account the fact that the fluxes of water and salt
are coupled. At steady state the water and salt must arrive
Results and Discussion
at the interface in the proportions required to form the
A. Sodium Sulfate Decahydrate. If a crystal retains
its shape as it grows, its mass (M) and surface area (S) are decahydrate. The salt flux carries with it water superim-
posed upon the diffusive salt flux (Sherwood et al, 1975).
related to its characteristic linear dimension, L, as The fate of salt transport is therefore greater than pre-
m =
PcAD (1) dicted when the coupling of the fluxes is neglected. In the
limit of a small difference between X; and Xb, Ns k[F/(F =

S = BL2 (2) -

Xb)] (CB Cj) > k(CB C,).


- -

The decahydrate crystal was approximated as a rectangular The effect of heat transport can be estimated by cal-
parallelopiped with A = 1.0 and B = 6.0. culating the temperature change at the interface from a
The change in crystal mass, m, was related to a change combined heat and mass balance
in the average linear dimension of the crystal, assuming
the rate of change of the crystal’s mass is proportional to h(Ti -

Th) =
k(-AH{)CtF In (5)
the exposed surface area
m(t)1/s =
m(t =
0)1/3 + (pcA)1^3(dL/dt)t (3) AH{ is the heat of fusion of the decahydrate. The ratio
of the mass to heat transport coefficients, k/h, requires
This equation, developed from eq 1, was first used by information about the coefficients independently. The
Hixson (1951) in analyzing the growth of crystals. ratio will be assumed to be that estimated from the
The results of the experiments are presented in Figure
“creeping flow” estimate to Brian and Hales (1969)
2, which shows the dependence of the rate of change of the
crystal dimension, L, on the concentration driving force k J_ /Pr\ / Sh\
=
~ =
"

defined in eq 4. The data were measured in the solution h p8Cp VSc/ \Nu)
temperature range 25.7 to 27.4 °C for growth rates and 23.8
/ Pr\ l12
to 28.6 °C for solution rates. Supersaturation ratios, (CB
CE)/CE, from -0.116 to +0.166 were used. A comparison
1

\Sc
[ 44 ++ 1.21(FeSc)2/3

[ 1.21 (RePr)V3
-

p8Cp )
of the rates of growth and solution shows that: (a) the rate J

of growth per unit driving force was slightly greater than Pr and Sc represent the Prandtl number and the Schmidt
the rate of solution; (b) both the rates of growth and so- number, respectively. Re is the Reynolds number based
lution were strong functions of the fluid velocity. on solution velocity and average particle diameter. Since
It appears that the rate of incorporation of molecules there is apparently no resistance due to surface kinetics
was not a rate-limiting factor in the velocity range studied. the interface composition, Xi; is set equal to the equilib-
Ind. Eng. Chem. Fundam., Vol. 23, No. 2, 1984 145

Figure 3. Dependence of crystal growth on fluid velocity and com-


parison with diffusion theory.

rium concentration Ce at the interface temperature.


The rate of change of the crystal dimension is related CONCENTRATION DRIVING FORCE
( g mole/1 iter )
to the flux by a mass balance at the crystal surface
Figure 4. Dependence of rate of growth and solution of anhydrous
^ =
(B/3A) (Nt/pm) (7) Na2S04 on concentration driving force.

pm is the crystal molar density. The mass transfer coef- excess at the crystal surface, C; Ce, required to produce
-

ficients extracted from the data in Figure 2 using eq 4 are the rate of growth of decahydrate observed in our exper-
presented in Figure 3, for two cases: (1) the interface iments. At a growth rate on the order of 1CT6 cm/s at 25.0
temperature was assumed equal to the bulk temperature °C the concentration driving force at the surface, Ci Ce, -

of the solution; (2) the interface temperature was predicted would be on the order of 10“6 g-mol/L, compared to an
by the combined heat and mass balance, eq 5 and 6. The observed overall driving force, Cb Ce, of 10"1 g-mol/L.
-

resulting mass transfer coefficients, expressed in dimen- The resistance introduced by this surface rate process is
sionless form, as Sherwood numbers, kL/D, were compared negligible compared to the mass transport resistance.
with values estimated from the creeping flow model of Data reported by McCabe and Stevens (1951) for copper
Brian and Hales (1969) in which the diffusivity was esti- sulfate pentahydrate and by Liu et al. (1971) for magne-
mated to be 5 X 1CT6 cm2/s and the crystal density pm was sium sulfate heptahydrate show appreciable surface re-
1.464 g/cm3. The following conclusions can be drawn: (a) sistance at low solution velocities. The lower rates might
The mass transfer coefficients were somewhat larger than be explained by dislocation theories of growth like that of
predicted by the creeping flow model. This is reasonable Burton, Cabrera, and Frank (1951). (1) Sodium sulfate
because the crystals are not spherical, (b) The mass is more heavily hydrated than either the copper or the
transfer coefficients computed from the rate of growth data magnesium hydrates. (2) The additional water molecules
were larger than those from the rate of solution experi- surrounding the ions weaken interionic forces in the
ments, possibly due to the effects of natural convection, crystal. (3) The crystal lattice is more easily deformed and
(c) Inclusion in the computations of the effect of heat dislocations are more easily created by stresses in the
transport led to a two to 12% increase in the mass transfer decahydrate. (4) Dislocations at the crystal surface act as
coefficients. sites for growth. The greater the dislocation density the
The density of the solution is a strong function of con- more likely a molecule at the surface will find a favorable
centration. Natural convection currents occur due to the site for attachment and the faster the crystal will grow.
concentration and temperature differences between the One therefore expects the crystals of decahydrate to grow
interface and the bulk solution. In the range of fluid at a faster rate than the other hydrates.
velocities studied the effect of natural convection may be B. Anhydrous Sodium Sulfate. The change in crystal
significant. During growth the fluid at the crystal surface mass was converted to a rate of change of the characteristic
is less dense than the bulk solution. The fluid rises past linear dimension, L, using eq 3. L is the distance from the
the surface, reinforcing the upward forced convection flow. crystal origin to the (111) face along the face-normal. The
During dissolution the fluid at the interface is denser than shape factors were assumed to be A = 9.32 and B = 34.19.
the bulk solution. Then the resulting natural convection The results of a series of rate-of-growth and solution
patterns act opposite to the forced convection, reducing experiments are shown in Figure 4. Driving forces were
the net upward flow. created by heating or subcooling a 3.04 M solution. The
Hayakawa et al. (1973) report data on the rate of growth results show that: (a) the effect of solution velocity on the
and solution of single crystals falling freely through a rate of growth is small compared to its effect on the rate
column of solution. From the rate of growth data, surface of dissolving; (b) the rate of growth is an order of mag-
concentrations and temperatures were calculated from nitude smaller than the rate of solution. These results
mass and heat transfer correlations developed from the indicate a significant resistance to growth at the surface,
rate of solution data. A surface rate equation was extracted compared to the mass-transport resistance in the solution
from the data based on the calculated interface properties. phase. A further set of experiments on the rate of growth
Their equation is at constant temperatures yielded the results shown in
d£ 6 R7 X 1 f)17 Figure 5. The data show a significant effect of temper-
df --exp(-27.1 X 103 RT) (C, -

C.) (8) ature.


Pm The experimental results were analyzed as follows to
This expression can be used to estimate the concentration determine the surface kinetics in the rate of growth pro-
146 Ind. Eng. Chem. Fundam,, Vol. 23, No. 2, 1984

(cb-c#)
{ g mole/liter )

Figure 5. Determination of activation energy for rate of growth of


anhydrous N2S04.
Figure 6. Electron micrograph of anhydrous crystal surface.
Table I
Sherwood no., kL/D
est from
creeping flow
soln vel, from eq of Brian and
cm/s expt Hales (1969)
8.2 39 64
35.0 117 100

cess: (a) The heat of solution of sodium sulfate is 0.28


kcal/g-mol. The effect of the heat generated on the tem-
perature at the interface is insignificant, so the effect of
heat transport can be ignored, (b) The rate of solution is
mass-transport controlled. Mass transfer coefficients were
extracted from the data with eq 4 and 7. The values are
listed in Table I.
The effect of the mass transfer resistance on the rate
of growth can be estimated by using the mass transfer
coefficients extracted from the rate of solution data to
calculate the interface concentration at different growth 7. Electron micrograph of anhydrous crystal surface.
rates. Rearranging eq 4 and 7
Figure

/ f PJ3A/B)dL/dt 1 \ placed in a desiccator. The dried crystal was gold coated

*b)^l-exP|--j] (9)
and viewed under an electron microscope at 640X mag-
nification. Both micrographs show a step-like array of
surface defects. Figure 7 contains a small crystallite of
sodium sulfate growing on the surface among these steps.
For rates of growth on the order of 10 6 cm/s at a fluid It is possible that the visible steps are actually “bunches”
velocity of 8.2 cm/s the mass transport resistance leads of much smaller steps of smaller height, where the actual
to a 1 % difference in concentration between the interface
process of addition of ions into the crystal lattice occurs.
and the bulk solution. A potential cause of the formation of “step-bunches” is the
The interface concentration was calculated for the data
presence of impurities on the crystal surface which impede
in Figure 5 by using eq 9. The rate of growth data were the motion of individual steps across the surface, causing
fit to the following empirical equation in terms of the the following steps to “bunch-up” behind the impeded step.
interface composition and the temperature The advance of these “bunches” of steps across the crystal
surface during the growth process has been described by

= 1.484 x 105 exp(-13.7 X W/RT)(C, -

Ce)16 (10)
dt Frank (1958) by use of kinematic wave theory.
It is impossible to say whether the visible features are
where Cj Ce is the concentration driving force at the
-

the product of the growth process alone or whether the


interface in g-mol/L. The activation energy for the process
steps were heavily etched during the cleaning process. In
is 13.7 kcal/g-mol compared to an activation energy of 3-4
any event, it is probable that the rate of growth of the (111)
kcal/g-mol for a diffusion-controlled process. There were crystal face is limited by the advance of such steps or
no other experimental data available for comparison with
“step-bunches” across the crystal surface.
these results.
Electron micrographs of a (111) face of an anhydrous Conclusions
crystal are shown in Figures 6 and 7. The crystal used The rates of growth and solution of single crystals of
for these pictures was taken from the growth cell during sodium sulfate decahydrate and anhydrous sodium sulfate
a typical growth experiment, dipped briefly in distilled were studied under conditions of low fluid velocity and low
water and placed in a container filled with petroleum ether. driving forces such as those likely to be encountered in a
The dried crystal was thoroughly rinsed in the ether and thermal energy storage process application. The following
Ind. Eng. Chem. Fundam. 1984, 23, 147-153 147

conclusions can be reached: (a) The rates of growth and Gr =


p2gL36c(AC)/ p2, Grashoff number
solution of sodium sulfate decahydrate are controlled by Nu =
Hl/k, Nusselt number
the rates of heat and mass transfer near the crystal surface, Pe =
(Re)(sc), Peclet number
(b) The rate of solution of anhydrous sodium sulfate is also PeT
=
(2r/D)(dr/dt), Peclet number due to radial velocity
controlled by the rate of mass transport from the crystal Re =
LVptjp, Reynolds number
surface, (c) The rate of growth of the (111) face of the
Sc =
pjpfi, Schmidt number
Sh =
kL/D, Sherwood number
anhydrous crystals is limited by the rate of incorporation
of the ions into the crystal lattice at the surface. Greek Letters
a =
fractional solution specific volume change
Acknowledgment dT temperature coefficient of fractional volume change, °C_1
=

dc = concentration coefficient of fractional volume change,


We gratefully acknowledge the support of the Depart-
(g-mol/L)'1
ment of Energy in the form of Research Grant No. DE- k =
solution thermal conductivity, cal/(cm s °C)
FG02-79ET-00088. The work reported here has been pc = crystal phase density, g/cm3
taken in part from the Master of Chemical Engineering Pm
=
crystal phase molar density, g-mol/cm3
Thesis (June, 1981) of David Rosenblatt. ps
= solution density, g/cm3
p = solution viscosity, g/(cm s), P
Nomenclature Subscripts
A = volume shape factor i = refers to interface properties
B =
surface area shape factor b = refers to bulk solution properties
C = concentration of sodium sulfate in solution, g-mol/cm3 e = refers to equilibrium properties
Ct = total concentration in solution, g-mol/cm3 Registry No. Sodium sulfte decahydrate, 7727-73-3; sodium
Cp
= heat capacity of the solution, cal/(g °C) sulfate, 7757-82-6.
D = diffusivity of sodium sulfate in solution, cm2/s
F = ratio of the salt flux to the total material flux Literature Cited
g = gravitational constant, cm/s2 Brian, P. L. T.; Hales, H. B. AIChEJ. 1969, 15, 419-425.
h = heat transport coefficient, cal/(cm2 s °C) Bruton, W.; Cabrera, H.; Frank, F. C. Phil. Trans R. Soc. London 1951,
AZ43, 299-358.
AHf = heat of fusion of sodium sulfate decahydrate, cal/g-mol Frank, F. C. “On the Kinematic Theory of Crystal Growth and Dissolution
AHs = heat of solution of anhydrous sodium sulfate, cal/g-mol Process” in “Growth and Perfection of Crystals", “International Crystal
k = mass transfer coefficient, cm/s Conference", Doremus, R.; Turnbull, J., Ed.; New York, Wiley: 1958; pp
411-418.
kj = crystallization rate coefficient, (cm/s)/(g-mol L)1-6 Hayakawa, T.; Matsuoka, M. Heat Transfer Jap. Res. 1973, 2, 104-115.
L = characteristic crystal dimension, cm Hixson, A. W.; Knox, K. L. Ind. Eng. Chem. 1951, 43, 2146.
m = crystal mass, g Liu, V. Y.; Tsuei, H. S.; Youngqulst, G. R. Chem. Eng. Prog. Symp. Ser.
N, = flux of salt to the interface, g-mol/(cm2 s) 1971, 110.
Nv = flux of water to the interface, g-mol/(cm2 s) McCabe, W. L.; Stevens, R. P. Chem. Eng. Prog. 1951, 47, 168-174.
r = equivalent crystal radius, cm
Sherwood, T. K.; Pigford, R. L.; Wilke, C. R. “Mass Transfer”; McGraw-Hill:
New York, 1975.
R = gas constant, cal/(g-mol K) Wetmore, F. E. W.; LeRoy, D. J. “Principles of Phase Equilibria”; McGraw-Hill:
T = temperature of the solution, K New York, 1951.
t = time, s
V = bulk solution velocity past the crystal, cm/s Received for review November 9, 1981
X = mole fraction of sodium sulfate in solution Accepted October 10, 1983

Interstitial Flow Intensification within Packed Granular Bed Filters:


Experiments and Theory
Robert W. L. Snaddon* and Peter W. Dietz
General Electric Company, Corporate Research and Development, Schenectady, New York 12301

The mechanical collection efficiency of a granular bed filter is determined by the flow within the bed. In spite of
the demonstrable sensitivity of existing models to their proposed flow structures, most do not attempt accurate
characterization of the flow. In particular, interstitial flow intensification and flow separation are usually neglected.
In the present paper an analytic model and collection measurements are combined to examine these effects. A
potential flow model is developed in which the magnitude of the flow intensification is characterized by a single
“intensification parameter". The experimental data are then employed to develop a unique correlation of this
parameter with Reynolds number.

Introduction of these devices as well as their ability to remove fine


In recent years considerable interest has been expressed submicron particulates, especially when augmenting
in the use of packed granular bed filters for separating fine agencies such as electrostatics are employed as shown by
particles and droplets from gas streams. This interest Kallio et al. (1979), Kallio and Dietz (1981), Dietz (1981),
stems from the inherent simplicity, reliability, and low cost Zahedi and Melcher (1976), and Zahedi and Melcher

0196-4313/84/1023-0147$01.50/0 © 1984 American Chemical Society

You might also like