Adsorption of Water and Ethanol
Adsorption of Water and Ethanol
Adsorption of Water and Ethanol
A pilot scale adsorber apparatus was designed and constructed to investigate water and ethanol adsorption/
desorption kinetics on 3A zeolite for the design purposes of a fuel ethanol dehydration pressure swing adsorption
(PSA) process. Equilibrium studies have shown that 3A zeolite adsorbed a significant amount of water while
very weak ethanol adsorption was observed. The breakthrough curves were utilized to study the effects of
column pressure, temperature, flow rate, pellet size, and adsorbate concentration on the overall mass transfer
resistance. Based on experimentally observed trends, both macropore and micropore diffusion were identified
as relevant mass transfer mechanisms. A mathematical model for a bench scale adsorption bed included the
linear driving force (LDF) adsorption rate model and the variation of axial velocity. A detailed heat transfer
model was a necessity since the bed dynamics was affected by heat transfer in the bed wall. The model was
used to analyze the experimental data and extract values of pertaining diffusion coefficients.
Table 1. Mathematical Model of Isobaric Nonisothermal Fixed Bed Adsorber with Initial and Boundary Conditions
in the bed wall, heat transfer through the bed wall from the The fluid phase axial thermal conductivity, λL, was estimated
zeolite bed, and heat transfer through the insulation to the by the correlation given by Dixon:17
environment. By considering such a detailed model, two
additional parameters were introducedsheat transfer coefficient 1 0.73εB 0.5
at the wall hw and the external heat transfer coefficient ) + (12)
PeAF Re Pr 9.7εB
heslumping the thermal resistance of the insulation and the 1+
Re Pr
thermal resistance due to the natural convection heat transfer
outside the adsorber.
The fluid film mass transfer coefficient was estimated from
The set of partial differential equations reported in Table 1
the correlation (13) given by Wakao and Funazkri.18 The
was solved by the method of lines using standard initial and
correlation is valid for a wide range of Reynolds numbers in
boundary conditions. Initially an adsorbate free bed was assumed
the range 3 < ReP < 104. The fluid-solid (particle) heat transfer
for adsorption, while the saturated bed initial conditions for gas
coefficient, hfs, was obtained by assuming an analogy between
and solid concentrations were used for the desorption modeling.
mass and heat transfer; the corresponding coefficient was
The discretization approximation in the axial direction by five
estimated from eq 14. The procedure was used with success in
point upwind finite differences15 resulted in the system of
the adsorption modeling at high Reynolds numbers.11
ordinary differential/algebraic equations (DAEs). Fifty axial
points were required to obtain a reasonable level of accuracy
and stability. The maximum time step of 1 s was used in all Sh ) 2.0 + 1.1Sc1/3Re0.6 (13)
calculations. The modeling of adsorption systems with a
nonlinear isotherm produces a set of stiff DAEs; hence, they Nu ) 2.0 + 1.1Pr1/3Re0.6 (14)
were solved by the DASSL code designed by Petzold16 using
an implicit backward differentiation formula. The FORTRAN Physical and transport properties such as gas viscosity,
IMSL Library subroutine DDASPG, based on DASSL, was used thermal conductivity, and specific heat capacity were calculated
to solve the problem in double precision arithmetic on a standard for a mixture of water or ethanol in the nitrogen carrier according
laptop 1.6 GHz computer. to generally accepted formulas and property correlations.19 All
important system characteristics are summarized in Table 2.
Estimation of Model Parameters The value of the volumetric heat source output QL in eq 6
The parameters in eqs 1-7 were either measured or deter- was evaluated from the initial conditions. Since the power output
mined from empirical correlations. The axial dispersion coef- of all band heaters was kept constant throughout the experiment
ficient, DL, was determined by the correlation suggested by and all other parameters were already known, the value for QL
Wakao and recommended by Ruthven:11 was estimated from the initial temperature profiles in the bed
with the aid of mathematical model.
DL The external heat transfer coefficient, he, was introduced by
20 1
) + (11) considering a rigorous energy balance for the adsorber wall.
udP Sc Re 2
The value of the coefficient he was determined experimentally
Ind. Eng. Chem. Res., Vol. 48, No. 20, 2009 9251
Table 2. Column and Adsorbent Characteristics
Column
adsorbent layer length, L 0.518 m
internal diameter, D1 0.034 m
outer diameter, D2 0.058 m
external diameter, D3 0.165 m
bed void fraction, εB 0.36
bulk density, FB 770 kg/m3
Wall (316 SS)
density, Fw 8000 kg/m3
heat capacity, cp,w 500 J · kg-1 · K-1
thermal conductivity, λw 17 W · m-1 · K-1
Adsorbent: W. R.Grace 3A Zeolite
pellet density, FP 1199.3 kg/m3
adsorbent particle radius, RP 1.785 × 10-3 m
pellet porosity, εP 0.37
BET surface area 45 m2/g
pellet thermal conductivity, λP 0.12 W · m-1 · K-1
pellet heat capacity, cp,s 1045 J · kg-1 · K-1
Figure 4. Equilibrium isotherms of water vapor on 3A zeolite. Data were
by heating the initially cold bed with the carrier gas (band measured at 100, 146, 167, and 200 °C for 3.6 mm pellets and at 167 °C
heaters were turned off) until steady state was reached. for 1.8 mm pellets. Solid lines represent the fit with the Langmuir isotherm
model.
Nuw ) 0.2Pr1/3Re0.8 (15) Table 3. Langmuir Equilibrium Model and Parameters for Water
on 3A Zeolite
Small radial temperature gradients in the bed (1-2 °C)
indicated that the major heat transfer resistance was in the b(T)Pw b∞ ) 5.3126 × 10-10 K0.5 Pa-1;
qw* ) qs,w(T) γ ) 23.235 ) Qst/R/T0; q0,w )
insulation and thus only an order of magnitude estimate for hw 1 + b(T)Pw
10.7446 mol/kg; δ ) 0.687 92;
was necessary. The value of the wall heat transfer coefficient, T0 ) 300 K
hw, did not have a significant effect on the generated temperature
curves. This was expected because of the large difference in
thermal conductivities of the steel wall and ceramic wool
b(T) )
b∞
√T
( )
exp γ
T0
T
insulation and turbulent flow conditions inside the bed. The
correlation given by Dixon, eq 15, was used to obtain an
estimate for the wall heat transfer coefficient.17 The developed
model was used to estimate the heat transfer coefficient for
((
qs,w(T) ) q0,w exp δ 1 -
T
T0 ))
various flows and temperatures. The values of the parameter he
showed only a weak dependence on the temperature and flow
rate, as can be seen from Figure 3. concentration, and after reaching the equilibrium, the adsorbate
uptake was evaluated through the overall mass balance.
Results and Discussion Figure 4 shows adsorption isotherms for water vapor obtained
at 100, 146, 167, and 200 °C on a 3A zeolite for particle diameters
Equilibrium Data. The adsorption breakthrough runs were 3.6 and 1.8 mm, respectively. The parameters of the equilibrium
used to obtain equilibrium data for adsorbing components. model were obtained by a nonlinear least-squares optimizer from
Initially clean bed was exposed to a step input of the adsorbate MATLAB. The Langmuir model with the temperature dependent
saturation capacity qs proved adequate. The values of the model
parameters are summarized in Table 3.
The isosteric heat of adsorption for the system water-3A
zeolite was evaluated from the value of parameter γ (see Table
3) to be 57.95 kJ/mol. Sowerby8 and Carmo20 used the values
50 and 43 kJ/mol. Lalik et al. have measured the heat of
adsorption of water on a 3A zeolite in the range 57-72 kJ/mol
in a microcalorimetric study.21 Gorbach’s work with a 4A zeolite
assumed 54.9 kJ/mol.10
The ethanol uptake by a 3A zeolite was studied at temper-
atures 100, 146, and 167 °C, respectively. The experimental
run carried out at 167 °C is shown in Figure 5. An immediate
breakthrough was observed, indicating a very low bed capacity
for ethanol.
Equilibrium loading of 0.03 mol/kg was obtained for the first
step in the experimental run E3. A very low ethanol loading
can be predicted from the temperature profiles as well. A slight
temperature increase can be observed for the first step only,
Figure 3. Comparison of experimental (points) and calculated (lines)
indicating that all sites available for ethanol adsorption were
temperature profiles used for estimation of external heat transfer coefficient occupied and no further adsorption takes place in the following
he for temperatures 100 and 167 °C and flows of 7.3, 15, and 30 SLM. steps. The fact that the small exothermic spike for the first step
9252 Ind. Eng. Chem. Res., Vol. 48, No. 20, 2009
Table 4. Experimental Conditions, Fitted kLDF Values, and Mass Transfer Resistances
resistance [s]
run no. P [kPa] T [°C] us [m/s] PH2O [kPa] FN2 [SLM] FH2O [g/h] dP [mm] 103kLDF [s-1] external film macropore micropore
0.078 8.2 15 13.5 3.57 0.90 51 376 685
w1 448 146
0.083 34.3 15 60 3.57 0.90 15 13 1084
0.083 8.2 15 13.5 3.57 0.55 73 605 1140
w2 448 100
0.083 23.5 15 40 3.57 0.55 N/A N/A N/A
0.088 8.2 15 13.5 3.57 8.50 18 100 0
w3 448 200
0.086 23.5 15 40 3.57 8.50 11 58 48
0.083 8.2 15 13.5 3.57 1.90 37 259 230
w4 448 167
0.082 23.5 15 40 3.57 1.90 17 67 442
0.083 8.2 7.35 13.5 3.57 2.60 26 130 229
w5 224 167
0.08 8.2 7.35 13.5 3.57 2.60 26 96 262
0.083 8.2 23.2 13.5 3.57 1.50 46 399 222
w6 689 167
0.082 8.2 23.2 13.5 3.57 1.50 46 297 324
0.083 12.3 15 20.5 3.57 2.00 28 202 269
w7 448 167
0.082 12.3 15 20.5 3.57 2.00 28 136 336
0.085 16.4 15 27.5 3.57 2.30 23 167 245
w8 448 167
0.082 16.4 15 27.5 3.57 2.30 23 102 310
0.167 8.2 30 27 3.57 2.20 26 259 169
w9 448 167
0.163 8.2 30 27 3.57 2.20 26 193 236
0.218 3.1 40 13.5 3.57 1.95 36 399 77
w10 448 167
0.221 8.2 40 36 3.57 1.90 22 193 311
0.083 8.2 15 13.5 1.785 9.5 12 65 28
w11 448 167
0.082 8.2 15 13.5 1.785 9.5 12 48 45
0.083 8.2 23.5 13.5 1.785 5.5 15 102 65
w12 689 167
0.082 8.2 23.5 13.5 1.785 5.8 15 74 83
0.083 12.3 15 20.5 1.785 10.5 9 51 35
w13 448 167
0.082 12.3 15 20.5 1.785 10.5 9 34 52
0.085 16.4 15 27.5 1.785 11.5 7 42 38
w14 448 167
0.082 16.4 15 27.5 1.785 11.5 7 25 54
0.167 8.2 30 27 1.785 8.5 9 68 41
w15 448 167
0.163 8.2 30 27 1.785 8.3 9 48 63
can be compared with the endothermic peak accompanying final A typical adsorption run is depicted in Figure 6, where the
desorption stage leads to the same conclusion. water breakthrough occurred after 1.25 h. The slope of the
The adsorbent selectivity for the ethanol-water mixture can breakthrough curve, especially in the early stage, is proportional
be now evaluated for the PSA feed stream conditionss92 wt to the adsorption rate. The steeper the curve the higher is the
% ethanol, 167 °C, 448.2 kPasusing the experimental value value of the mass transfer coefficient. An infinite value of the
for the ethanol and water equilibrium model from Table 3. The mass transfer coefficient would lead to a shock front and
value of ∼900 for the adsorbent selectivity (analogue to relative maximum bed utilization (dashed line in Figure 6). Mass transfer
volatility in distillation) illustrates the molecular sieving mech- has thus a negative dispersive effect on the column performance.
anism for ethanol-water separation by the PSA process. The shape of the concentration breakthrough curve is point-
Water Adsorption Kinetics Study. Breakthrough experiments symmetric to the center of the front for an isothermal system
were carried out to study the effect of the temperature, pressure, with a linear equilibrium isotherm; both conditions are strongly
water concentration, flow rate, and pellet size on the bed perfor- violated in the case of water adsorption on a 3A zeolite. The
mance and on the adsorption/desorption kinetics. The experimental effect of the isotherm shape was positive for adsorption since
apparatus and procedure were discussed earlier. Adsorption and for a favorable isotherm a fixed bed transition approaches a
desorption steps were performed in sequence for each experimental
run. Table 4 summarizes all performed experiments; here conditions
for adsorption are always shown first.
kT
λ) (16)
Pπdmolecule2√2
Now, if the mean free path is larger than the pore diameter
the Knudsen diffusion is the governing mechanism; if the pore
diameter is much bigger than λ the bulk diffusion regime
dominates. Consequently, a transition from bulk diffusion to
Knudsen diffusion is expected as the pressure decreases since
the mean free path increases. However, this transition is
observed only for diffusion in small pores or for diffusion of
larger molecules. If this was the case here, we would expect a
different effect of pressure on the slope of a breakthrough curve.
The Knudsen diffusion coefficient itself is not a function of
pressure, and thus small or no effect would be observed.
The effect of pressure on the adsorption kinetics was studied
Figure 9. Temperature and concentration profiles for water desorption on for both 3.6 and 1.8 mm pellets, and the same trends were
3A zeolite for bed temperatures of 200, 167, and 100 °C, respectively. observed. It is therefore concluded that the macropores make a
Temperatures were measured at four different locations: the bed inlet, L/3, significant contribution to the overall mass transfer resistance
2/3L, and the bed end.
and that macropore diffusion is controlled by the molecular
diffusion mechanism alone.
For desorption runs (see Figure 11) the effect of pressure on
the rate of adsorbent regeneration was not as significant as for
the adsorption. It seems that the desorption rate might be
controlled by other factors such as diffusion in micropores,
thermal effects, and isotherm effects.
Effect of Water Concentration. The effect of water partial
pressure in the feed stream was studied at temperature 167 °C,
pressure 448 kPa, and nitrogen flow rate 15 SLM for two
different pellet sizes; see experimental runs w4, w7, w8, w11,
w13, and w14. In both cases, water partial pressures studied
were 8.2, 12.3, and 16.4 kPa.
The stoichiometric breakthrough time θs decreased as the
water partial pressure increased; see Figure 12. This was
expected because the bed is saturated faster if there is more
water in the feed stream. Experimental breakthrough times for
Figure 10. Temperature and concentration profiles for water adsorption on 8.2 and 16.4 kPa were approximately 0.65 and 1.2 h, respec-
3A zeolite for pressures 224, 448, and 689 kPa, respectively. Temperature
tively. The corresponding equilibrium loadings q* were 4.76
at axial position of 2/3L is shown. Pellet size was 3.6 mm.
and 5.91 mol/kg, respectively. The values of the stoichiometric
breakthrough time θs were evaluated for experimental runs w4
167 °C; however, for the experiment at 100 °C water was still and w8 to 2.33 and 1.4 h, respectively. As mentioned earlier,
present even after 6 h of regeneration. θs corresponds to a shock breakthrough curve in the absence of
Effect of Pressure. The individual mass transfer resistances any mass transfer resistance; the ratio of θb/θs ) 1. It follows
for an adsorbent with bidisperse pore structure such as zeolite that the closer the breakthrough time ratio to a unity is the
are external laminar film, macropore resistance, and diffusion smaller the mass transfer resistance is. By inspection of the
in the zeolite crystals (micropores). Among these only the rate experimental and stoichiometric breakthrough times it is evident
of diffusion in macropores is affected by the pressure. Experi- that the system approaches infinite mass transfer as the feed
mental runs w4, w5, and w6 were carried out to confirm or concentration increases. Hence, it was concluded that, with
contradict the significance of the macropore resistance. Figure increasing partial pressures, sharper concentration profiles were
10 depicts the measured concentration and temperature profiles. observed as a consequence of the favorable isotherm affecting
the kinetic parameters. The same information can be obtained
It is evident that the slope of the breakthrough curve increased
from the temperature profiles, where an increase in the water
as the pressure decreased. An increase of the rate of adsorption
concentration leads to higher and steeper temperature curves.
can be realized from the temperature profiles as well. The hot
At the end of adsorption, the desorption step that followed
spot increased for lower pressures. It follows that the mass
maintained the same operating conditions; see Figure 13. The
transfer into the pellet must be affected by the molecular
slow process was due to the fact that desorption was controlled
diffusion mechanism since the mass transfer rate is inversely
by the unfavorable equilibrium isotherm and the situation got
proportional to the pressure. The same dependence on the worse as the concentration decreased.
pressure is observed for the molecular (bulk) diffusion coef-
Effect of Carrier Gas Flow Rate. Experimental runs w11
ficient in the Chapman-Engskog or Fuller equation.19 and w15 were designated to show the effect of the carrier gas flow
The diffusion mechanism, where the interactions molecule- rate on the adsorption and desorption performance. Experimental
wall are more frequent than the interactions among molecules, is conditions are summarized in Table 4. Measured concentration and
known as the Knudsen diffusion. From the kinetic theory of gases, temperature profiles for adsorption runs are depicted in Figure 14.
the mean free path (λ), the distance a molecule travels between Concentration and temperature profiles for run w15 looked almost
two collisions, is given by the following expression (16):5 identical to those from run w11. An increase of the flow rate had
Ind. Eng. Chem. Res., Vol. 48, No. 20, 2009 9255
Figure 11. Temperature and concentration profiles for water desorption on Figure 14. Adsorption runs w11 and w15. Temperature readings at axial
3A zeolite for pressures 224 and 448 kPa, respectively. Temperature at position z ) 2/3L and concentration profiles for carrier gas flow rate of 15
axial position of 2/3L is shown. Pellet size was 3.6 mm. and 30 SLM, respectively. Experimental conditions: 167 °C, 448 kPa, 8.2
kPa water partial pressure, and pellet size 1.8 mm.
The results for desorption runs w11 and w15 are shown in Figure
15. The regeneration time was halved when the flow rate was
doubled. The breakthrough time for adsorption with flow rate 15
SLM was less than 2 h; see Figure 14. At the same conditions the
regeneration took almost 6 h; however, when the purge flow rate
was doubled the regeneration process took only 3 h.
Effect of Particle Size. The size of the pellet affects many
system parameters such as film coefficients, dispersion in packed
bed, bed porosity, and intraparticle transport processes. The
effect of the first three is usually not significant in an industrial
adsorber; however, the last one can be detrimental if the
macropore diffusion is controlling the mass transfer rate. In other
words, the change of the pellet size will have no effect on the
breakthrough curve if the macropore diffusion is negligible.
Figure 13. Temperature readings at axial position z ) 2/3L and concentra- Experimental runs w9 and w15 were designated to address
tion profiles for water partial pressures 8.2, 12.3, and 16.4 kPa, respectively. this issue. Concentration and temperature profiles are plotted
Experimental conditions for desorption runs: 167 °C, 448 kPa, 15 SLM N2
flow, and pellet size 3.6 mm. in Figure 16. The breakthrough times for 3.6 and 1.8 mm pellets
were 0.44 and 0.89 h, respectively. This significant improvement
of the adsorber performance was due to the fact that the
a positive effect on the film mass transfer and heat transfer macropore resistance decreased by a factor of 4 since the
coefficients. The Reynolds number increased from 27 to 54. The diffusion time scale for macropores depends on the square of
profiles at a higher flow rate were slightly steeper, suggesting that the particle radius. This kinetic effect can be followed from the
film resistance was present but its contribution to the overall mass temperature profiles as well. The temperature peaks were steeper
transfer resistance was only minimal. and narrower due to an increase in the mass transfer rate.
9256 Ind. Eng. Chem. Res., Vol. 48, No. 20, 2009
Figure 19. Transient adsorber bed and wall temperature profiles as a function
of axial coordinate. Particular time values in hours can be found in the
legend. Calculated data correspond to the fit for experimental run w4
depicted in Figure 18, bottom.
(11) Ruthven, D. M. Principles of Adsorption and Adsorption Processes; (20) Carmo, M. J.; Gubulin, J. C. Ethanol-water adsorption on com-
Wiley-Interscience: New York, 1984. mercial 3A zeolites: kinetic and thermodynamic data. Braz. J. Chem. Eng.
(12) Simo, M.; Brown, C. J.; Hlavacek, V. Simulation of pressure swing 1997, 14 (3), 217–224.
adsorption in fuel ethanol production process. Comput. Chem. Eng. 2008, (21) Lalik, E.; Mirek, R.; Rakoczy, J.; Groszek, A. Microcalorimetric
32 (7), 1635–1649. study of sorption of water and ethanol in zeolites 3A and 5A. Catal. Today
(13) Glueckauf, E.; Coates, J. E. Theory of chromatography. J. Chem. 2006, 114 (2-3), 242–247.
Soc. 1947, 1315. (22) Ruthven, D. M.; Farooq, S.; Knaebel, K. S. Pressure Swing
(14) Sircar, S.; Hufton, J. R. Why Does the Linear Driving Force Model Adsorption; VCH Publishers, Inc.: New York, 1994.
for Adsorption Kinetics Work. Adsorption 2000, 6, 137–147. (23) Silva, J. A. C.; Rodrigues, A. E. Fixed-Bed Adsorption of n-Pentane/
(15) Schiesser, W. E. The Numerical Method of Lines, 1st ed.; Academic Isopentane Mixtures in Pellets of 5A Zeolite. Ind. Eng. Chem. Res. 1997,
Press: San Diego, CA, 1991. 36 (9), 3769–3777.
(16) Petzold, L. A description of DASSL: a differential algebraic system
solver. In IMACS Trans. Scientific Computing Vol. 1; Stepleman, R. S., (24) Do, D. D. Analysis of Adsorption Kinetics in a Zeolite Particle. In
Ed.; North-Holland: Amsterdam, 1993; pp 65-68. Adsorption Analysis: Equilibria and Kinetics; Imperial College Press: London,
(17) Dixon, A. G.; Cresswell, D. L. Theoretical prediction of effective 1998.
heat transfer parameters in packed beds. AIChE J. 1979, 25 (4), 663–676.
(18) Wakao, N.; Funazkri, T. Effect of fluid dispersion coefficients on ReceiVed for reView March 18, 2009
particle-to-fluid mass transfer coefficients in packed beds. Correlation of ReVised manuscript receiVed August 14, 2009
Sherwood numbers. Chem. Eng. Sci. 1978, 33 (10), 1375–1384. Accepted August 28, 2009
(19) Reid, R. C.; Prausnitz, J. M.; Poling, B. E. Properties of Gases
and Liquids, 4th ed.; McGraw-Hill: New York, 1987. IE900446V
ABOUT THERMAL KINETICS
At the forefront of new techniques and technologies, Thermal Kinetics has
been providing full-service process equipment engineering, development,
and design services since 1999. We offer advanced energy-saving
solutions, patented technological integration, and sophisticated process
plant development — proudly matching each of our customers’ specific
needs with optimum productivity and profitability.