You are on page 1of 11

4346

Ind. Eng. Chem. Res. 2000, 39, 4346-4356

Absorption of Carbon Dioxide into Aqueous Blends of


Diethanolamine and Methyldiethanolamine
Edward B. Rinker, Sami S. Ashour, and Orville C. Sandall*
Department of Chemical Engineering, University of California, Santa Barbara, California 93106-5080

In this work, a comprehensive model is developed for the absorption of carbon dioxide into
aqueous mixtures of primary or secondary alkanolamines with tertiary alkanolamines. The
model, which is based on penetration theory, incorporates an extensive set of important reversible
reactions and takes into account the coupling between chemical equilibrium, mass transfer,
and chemical kinetics. The reaction between CO2 and the primary or secondary amine is modeled
according to the zwitterion mechanism. The key physicochemical properties that are needed for
the model are the CO2 physical solubility, the CO2 and amine diffusion coefficients, and the
reaction rate coefficients and equilibrium constants. Data for carbon dioxide absorption into
aqueous mixtures of diethanolamine and methyldiethanolamine are compared to model
predictions.
Introduction
Removal of acid gas impurities, such as CO2, is
important in natural gas processing. Natural gas,
depending on its source, can have varying concentrations of CO2. Some of the CO2 is often removed from
natural gas because, when present at high levels, it
reduces the heating value of the gas, and because it is
costly to pump this extra volume when it does not have
any heating value.
Aqueous solutions of alkanolamines are the most
widely used solvents for removing CO2. The most commonly used alkanolamines are the primary amine
monoethanolamine (MEA), the secondary amine diethanolamine (DEA), and the tertiary amine methyldiethanolamine (MDEA). Primary and secondary amines react
rapidly with CO2 to form carbamates. By the addition
of a primary or secondary amine to a purely physical
solvent such as water, the CO2 absorption capacity and
rate is enhanced manyfold. However, because there is
a relatively high heat of absorption associated with the
formation of carbamate ions, the cost of regenerating
primary and secondary amines is high. Primary and
secondary amines also have the disadvantage of requiring 2 mol of amine to react with 1 mol of CO2; thus,
their loadings are limited to 0.5 mol of CO2/mol of
amine. Tertiary amines lack the N-H bond required to
form the carbamate ion and therefore do not react
directly with CO2. However, in aqueous solutions,
tertiary amines promote the hydrolysis of CO2 to form
bicarbonate and the protonated amine. Amine-promoted
hydrolysis reactions are much slower than the direct
reactions of primary and secondary amines with CO2,
and therefore, the kinetic selectivity of tertiary amines
toward CO2 is poor. However, the heat of reaction
associated with the formation of bicarbonate ions is
much lower than that associated with carbamate formation, and thus, the regeneration costs are lower for
tertiary amines than for primary and secondary amines.
Another advantage with tertiary amines is that the
stoichiometry is 1:1, which allows for very high equilibrium CO2 loadings.
* Corresponding author: sandall@engineering.ucsb.edu.

Chakravarty et al.1 suggested that, by mixing a


primary or secondary amine with a tertiary amine, the
CO2 selectivity in the presence of H2S could be improved
and regeneration costs minimized. These blended amine
solutions also offer the advantage of setting the selectivity of the solvent toward CO2 by judiciously mixing the
amines in varying proportions, which results in an
additional degree of freedom for achieving the desired
separation for a given gas mixture. This approach could
dramatically reduce capital and operating costs while
providing more flexibility in achieving specific purity
requirements. Because of the need to exploit poorer
quality crude and natural gas coupled with increasingly
strict environmental regulations, highly economical and
selective acid gas treatment is more important today
than at any time in the past. As a result, there has been
a resurgence of interest in improved alkanolamine solvents and particularly in aqueous blends of alkanolamines.
Design methods for acid-gas-treating processes employing aqueous blends of alkanolamines vary widely
in their effectiveness at predicting process performance.
Many acid-gas-treating processes are still designed by
experience and heuristics, resulting in overdesign,
excessive energy consumption, and often failure to meet
purity requirements entirely (Chakravarty et al.1).
Another common method uses equilibrium stage models
modified by tray efficiencies. This method, however,
requires the use of existing plant data and lumps all
uncertainties about the finite reaction rates of the gases
in the solvent into one parameter, the tray efficiency.
Such a model cannot be predictive and will not capture
the essential interplay of mass transfer, chemical kinetics, and chemical thermodynamics that occur in complex
chemical solvents such as aqueous blends of alkanolamines. The third method of design is to develop models
based on the chemistry and physics of the process, which
accounts for rates of mass transfer coupled with chemical kinetics and thermodynamics. These models, while
still requiring some experimental hydrodynamic information specific to different types of contacting devices,
are capable of predicting column performance, thus
minimizing the costs of design, equipment, and energy
consumption.

10.1021/ie990850r CCC: $19.00 2000 American Chemical Society


Published on Web 09/27/2000

Ind. Eng. Chem. Res., Vol. 39, No. 11, 2000 4347

The objective of the work presented here is to develop


a comprehensive model for the absorption of CO2 into
aqueous blends of tertiary and primary or secondary
amines. Experiments to test the model using methyldiethanolamine (MDEA) as the tertiary amine and diethanolamine (DEA) as the secondary amine are described.

Reaction Mechanism
When CO2 is absorbed into an aqueous solution of a
tertiary alkanolamine, R(1)R(2)R(3)N, and a primary or a
secondary alkanolamine, R(4)R(5)NH, the following reactions may occur:
K1,k1

CO2 + H2O 798 H2CO3

(1)

K2,k2

CO2 + OH- 798 HCO-

(2)

K15

HCO3- + H3O+ 798 H2CO3 + H2O

(15)

K16

2H2O 798 OH- + H3O+

(16)

For MDEA, we have R(1) ) -CH3 and R(2) ) R(3) )


-CH2CH2OH, and for DEA, we have R(4) ) R(5) -CH2CH2OH. Ki, ki, and k-i are the equilibrium constant, the
forward rate coefficient, and the reverse rate coefficient
for reaction i. Reactions 1-11 are considered to be
reversible with finite reaction rates. Whereas reactions
12-16 are considered to be reversible and instantaneous
with respect to mass transfer and at equilibrium, since
they involve only proton transfers.
Note that not all of the reaction equilibrium constants
are independent. Only eight equilibrium constants (K2,
K4, K5, K12, K13, K14, K15, and K16) are independent. The
remaining eight can be obtained by appropriate combinations of the independent equilibrium constants.
The interaction between the protonated and unprotonated amines according to the reaction

K3,k3

CO2 + R(1)R(2)R(3)N + H2O 798

K17

R(1)R(2)R(3)NH+ + HCO3- (3)


(4)

(5)

K4,k4,k-4

(4)

(5)

CO2 + R R NH 798 R R NH COO

(4)

K5,k5,k-5

R(4)R(5)NH+COO- + R(4)R(5)NH 798


R(4)R(5)NH2+ + R(4)R(5)NCOO- (5)
K6,k6,k-6

R(4)R(5)NH+COO- + R(1)R(2)R(3)N 798


R(1)R(2)R(3)NH+ + R(4)R(5)NCOO- (6)
K7,k7,k-7

R(4)R(5)NH+COO- + H2O 798


H3O + R(4)R(5)NCOO- (7)
K8,k8,k-8

R(4)R(5)NH+COO- + OH- 798


H2O + R(4)R(5)NCOO- (8)
K9,k9,k-9

R(4)R(5)NH+COO- + HCO3- 798


H2CO3 + R(4)R(5)NCOO- (9)
K10,k10,k-10

R(4)R(5)NH+COO- + CO32- 798


HCO3- + R(4)R(5)NCOO- (10)
K11,k11

R(4)R(5)NCOO- + H2O 798


R(4)R(5)NH + HCO3- (11)
K12

R(1)R(2)R(3)NH+ + OH- 798 R(1)R(2)R(3)N + H2O (12)


K13

R(4)R(5)NHH2+ + OH- 798 R(4)R(5)NH + H2O (13)


K14

HCO3- + OH- 798 CO32- + H2O

(14)

R(4)R(5)NH2+ + R(1)R(2)R(3)N 798


R(4)R(5)NH + R(1)R(2)R(3)NH+ (17)
involves only a proton transfer and is considered to be
instantaneous and at equilibrium. Reaction 17 is implicitly included in the reaction scheme above, as it can
be obtained by properly combining the instantaneous
equilibria reactions 12 and 13. Hence, we have K17 )
K13/K12.
The proposed mechanism for the reaction between
CO2 and tertiary alkanolamines, R(1)R(2)R(3)N, indicates
that they do not react directly with CO2. Instead,
tertiary alkanolamines act as bases that catalyze the
hydration of CO2 (Donaldson and Nguyen,2 Haimour et
al.,3 Versteeg and van Swaaij,4 Littel et al.,5 Rinker et
al.6). In contrast, the proposed mechanism for the
reaction between CO2 and a primary or secondary
alkanolamine, R(4)R(5)NH, involves the formation of a
zwitterion, R(4)R(5)NH + COO- (see reaction 4), followed
by the deprotonation of the zwitterion by a base to
produce carbamate, R(4)R(5)NCOO-, and protonated base
(see reactions 5-10) (Caplow,7 Danckwerts,8 Blauwhoff
et al.,9 Versteeg and van Swaaij,10 Versteeg and
Oyevaar,11 Versteeg et al.,12 Glasscock et al.,13 Littel et
al.,14 Rinker et al.15). Any base present in the solution
might contribute to the deprotonation of the zwitterion.
The contribution of each base would depend on its
concentration as well as its strength. Hence, the main
contribution to the deprotonation of the zwitterion in
an aqueous solution of a mixture of a primary or
secondary alkanolamine, R(4)R(5)NH, and a tertiary
alkanolamine, R(1)R(2)R(3)N, would come from R(4)R(5)NH,
R(1)R(2)R(3)N, and to a lesser extent OH- and H2O. There
are two limiting cases in the zwitterion mechanism.
When the zwitterion formation reaction is rate-limiting,
the reaction rate appears to be first-order in both the
amine and CO2 concentrations. In the case of monoethanolamine (MEA), a primary alkanolamine, the
formation of the zwitterion has been shown to be the
rate-determining step (Danckwerts,8 Sada et al.,16 Versteeg and van Swaaij,10 Littel et al.14). On the other
hand, when the zwitterion deprotonation reactions are

4348

Ind. Eng. Chem. Res., Vol. 39, No. 11, 2000

rate-limiting, the overall reaction rate appears to be of


order 2 in the amine concentration. For the secondary
alkanolamines diethanolamine (DEA) and diisopropanolamine (DIPA), the rate-determining step is the deprotonation of the zwitterion (Littel et al.13). Several
authors have reported some rate coefficients for this
limiting case of the zwitterion mechanism for DEA and
DIPA for a few temperatures (Blauwhoff et al.,9 Versteeg and van Swaaij,16 Versteeg and Oyevaar,10 Glasscock et al.,12 Rinker et al.14). Similarly, if neither
reaction in the zwitterion mechanism is rate-limiting,
the reaction rate exhibits a fractional order between 1
and 2 with respect to the amine concentration. However,
the rate expression is more complicated than the limiting cases. Fractional orders are usually only observed
for reactions between CO2 and secondary amines (Sada
et al.,16 Versteeg and Oyevaar,11 Littel et al.14); however
fractional orders have also been observed for sterically
hindered primary amines such as 2-amino-2-methylpropanol (AMP) (Bosch et al.,17 Alper18).

All reactions at equilibrium


(only independent equilibrium constants)
K2 )

u04

K4K5K13K16 )
K12 )
K13 )
K14 )

u07 u010

u02
u08
u06

u2 ) [R(1)R(2)R(3)N]
-

u4 ) [HCO3 ]

u3 ) [R R R NH ]
u5 ) [OH-]

u6 ) [CO32-]

u7 ) [H3O+]

u8 ) [R(4)R(5)NH]

u9 ) [R(4)R(5)NH2+]
u11 ) [H2CO3]

u10 ) [R(4)R(5)NCOO-]
u12 ) [H2O]

Liquid Bulk Concentrations of All Chemical


Species. The liquid bulk concentrations of all chemical
species can be estimated from the initial concentrations
of R(1)R(2)R(3)N and R(4)R(5)NH; the initial CO2 loading
of the solution, L1; and the assumption that all reactions
are at equilibrium. Because the concentration of water
is much larger than the concentrations of all other
chemical species, changes to its concentration over very
short contact times are negligible, and we assume that
its concentration remains constant. Hence, we only need
to solve for the concentrations of the remaining 11
chemical species. We have the following equations for
0
the liquid bulk concentrations u01, ... , u11
:

u011

u03

(1)

(2)

(3)

)[R R R N]initial

R1 ) -k1u1 +

(18)

(19)

R3 ) -k3u1u5 +

u01 + u04 + u06 + u010 + u011 + L1{[R(1)R(2)R(3)N]initial +

R4,...,10 )

[R R NH]initial} (20)
(5)

Electroneutrality balance
u03 + u07 + u09 - u04 - u05 - 2u06 - u010 ) 0

(21)

k1
u
K1 11

where

(29)

k2
u
K2 4

(30)

k3
uu
K3 3 4

(31)

R2 ) -k2u1u5 +

Overall carbon (from CO2) balance


(4)

(28)

0
) and 11 nonlinear
We have 11 unknowns (u01, ... , u11
algebraic equations that we can solve for the liquid
bulk concentrations. We have found that Newtons
method did not converge unless the initial guesses
for the liquid bulk concentrations were very close to
the (unknown) solution. We, therefore, used the
Newton homotopy continuation method (Hanna and
Sandall19), which exhibited better convergence behavior.
The Partial Differential and Nonlinear Algebraic Equations That Describe the Diffusion/
Reaction Processes. Higbies penetration model (Higbie,20 Danckwerts21) was used to set up the diffusion/
reaction partial differential equations that describe the
absorption/desorption of CO2 into/from aqueous solutions of tertiary amines, R(1)R(2)R(3)N, and primary or
secondary amines, R(4)R(5)NH, in a laminar-liquid-jet
absorber or a stirred-cell absorber. All reactions were
treated as reversible reactions. The first 11 reactions
have finite reaction rates, which are given by the
following reaction rate expressions, where Ri is the
reaction rate expression for reaction i:

Overall primary or secondary amine, R(4)R(5)NH,


balance
u08 + u09 + u010 )[R(4)R(5)NH]initial

(27)

u04 u07

Overall tertiary amine, R(1)R(2)R(3)N, balance


u02

(26)

u04 u05

K16 ) u05 u07

(3)

(25)

u09 u05

For convenience, the concentrations of the chemical


species are renamed as follows:

(2)

(24)

u03 u05

K15 )

(1)

(23)

u01 u08

Mathematical Model

u1 ) [CO2]

(22)

u01 u05

-k4 u1u8 1+

1
B

(BA)u ]
10

(32)

Ind. Eng. Chem. Res., Vol. 39, No. 11, 2000 4349

( )

A)

( ) ( )
( ) ( ) ( )

k5 u 9
+
k-4 K4K5
k8
k-4

k6 u3
+
k-4 K4K6
u12
k9
+
K4K8 k-4

k7 u7
+
k-4 K4K7
u11
k10 u4
+
(33)
K4K9 k-4 K4K10

and

B)

( ) ( ) ( ) ( )
( ) ( )

k5
k6
k7
k8
u +
u +
u +
u +
k-4 8
k-4 2
k-4 12
k-4 5
k9
k10
u +
u (34)
k-4 4
k-4 6
R11 ) -k11u10 +

k11
uu
K11 4 8

(35)

Note that eq 32 was derived with the assumption of


a pseudo-steady-state approximation for the zwitterion
reaction intermediate, R(4)R(5)NH+COO-. The partial
differential equations that describe the diffusion/reaction processes were combined so as to eliminate the very
large reaction rates for the instantaneous reactions 1216. Because these reactions are assumed to be at
equilibrium, their equilibrium constant expressions
were used to complete the equations that are required
to solve for the concentration profiles of all chemical
species. Furthermore, the diffusion coefficients of the
ionic species were assumed to be equal. With this
assumption, the electrostatic potential gradient terms
in the diffusion/reaction partial differential equations
for the ionic species can be neglected, while the electroneutrality of the solution is preserved. The more
rigorous approach of taking into account the electrostatic potential gradient terms with unequal diffusion
coefficients for the ionic species requires much greater
computational effort with essentially intangible effects
on the predicted rates of absorption. The following
equations describe the diffusion/reaction processes:

u10
2u10
) D10 2 + R11 - R4,...10
t
x

(41)

Instantaneous reactions assumed to be at equilibrium

K12 )

u2
u3u5

(42)

K13 )

u8
u9u5

(43)

K14 )

u6
u4u5

(44)

K15 )

u11
u4u7

(45)

K16 ) u5u7

(46)

We have 11 partial differential/algebraic equations that


we can solve for the concentrations of the 11 chemical
species, u1, ... , u11.
Initial Condition and Boundary Condition at x ) .
At t ) 0 (for all x ) 0) and at x ) 8 (for all t ) 0), the
concentrations of all chemical species are equal to their
liquid bulk concentrations.

ui ) u0i , i ) 1, ... , 11

ui
) 0 at x ) 0, t > 0
x

2u1
2u4
u1u4 u6 u10 u11
+
+
+
) D1 2 + D4 2 +
t t
t
t
t
x
x
2u6
2u10
2u11
D6 2 + D10 2 + D11 2 (37)
x
x
x
Total tertiary amine, R(1)R(2)R(3)N, balance
(38)

Total primary or secondary amine, R(4)R(5)N,


balance
u8 u9 u10
2u8
2 u 9
2u10
+
+
) D8 2 + D9 2 + D10 2
t
t
t
x
x
x

Carbamate, R(4)R(5)NCOO-, balance

(47)

(36)

Total carbon (from CO2) balance

u2 u3
2 u 2
2 u 3
+
) D2 2 + D 3 2
t
t
x
x

u3 u7 u9 u4 u5
u6 u10
+
+
-2
)
t
t
t
t
t
t
t
2u3
2u7
2 u 9
2 u 4
2u5
D3 2 + D7 2 + D9 2 - D4 2 - D5 2 x
x
x
x
x
2
2
u6
u10
2D6 2 - D10 2 (40)
x
x

Boundary Condition at Gas-Liquid Interface (x ) 0).


At x ) 0 (gas-liquid interface), the fluxes of the
nonvolatile chemical species are equal to zero, which
leads to the following equations:

CO2 balance
2u1
u1
) D1 2 + R1 + R2 + R3 + R4,...10
t
x

Electroneutrality balance

(39)

(48)

for all i except i ) 1 (CO2). For the volatile component,


CO2, the mass transfer rate in the gas near the interface
is equal to the mass transfer rate in the liquid near the
interface.

u1
) kg,1[P1 - H1u1(0,t)] at x ) 0, t > 0
-D1
x

(49)

H1 is the physical equilibrium constant (Henrys law


constant) of CO2, which is defined as the interfacial
partial pressure of CO2 in the gas phase, P/1, divided by
the interfacial concentration of CO2 in the liquid, u/1.
For the case of pure CO2 in the gas phase, the interfacial
partial pressure of CO2, P/1, is the same as the bulk
partial pressure of CO2, P1, and there is not any mass
transfer resistance in the gas-phase (kg,1 f ). Hence,

4350

Ind. Eng. Chem. Res., Vol. 39, No. 11, 2000

the boundary condition for CO2 at the gas-liquid


interface reduces to

u1(0,t) ) u/1 )

P1
at x ) 0, t > 0
H1

d2l
4Q

D1

D1

u1
(0,t) dt
x

(52)

RA1
0
kl,1
(u/1- u01)

Table 1 gives the values of the equilibrium constants


used in the model calculations, with the exception of
K15. The value of K15 at 298 K was taken to be 2 10-4
m3/kmol (Cotton and Wilkinson31). K15 was then corrected for temperature dependence according to the
following equation:

K15(T)

K15(298.15)

1
-H0 1
R T 298.15

(57)

where the standard enthalpy change of reaction, H0,


is assumed to be independent of T and is approximated
0
(Smith and Van
by its value at 298.15 K, H298.15
0
Ness32). H298.15 values were calculated from values
reported in the CRC Handbook of Chemistry and
Physics (Lide33).
Experimental Apparatus and Procedure

(53)

The enhancement factor of CO2 is determined according to the following equation:

E1 )

(56)

ln

The average rate of absorption of CO2 per unit interfacial area is then computed from the following equation:

RA1 ) -

2895
T

log10(k2) ) 13.635 -

(51)

where Q is the volumetric liquid flow rate, and d and l


are the diameter and length of the laminar-liquid jet,
respectively. For Higbies penetration model, the liquidphase mass transfer coefficient for physical absorption
of CO2 is defined as
0
)2
kl,1

17265.4
(55)
T

(50)

The differential equations are integrated from t ) 0


to t ) , the contact time. For the laminar-liquid-jet
absorber, the contact time is given by

log10(k1) ) 329.85 - 110.54 log10(T) -

(54)

where u/1 and u01 are the interfacial and bulk concentrations of CO2 in the liquid, respectively.
Method of Solution for the Partial Differential/
Algebraic Equations. The method of lines was used
to transform each partial differential equation into a
system of ordinary differential equations in t (Hanna
and Sandall19). The systems of partial differential/
algebraic equations were transformed into larger systems of ordinary differential/algebraic equations, which
were then solved by using the code DDASSL (Petzold22)
in double-precision Fortran on an HP-735 computer.
Model Parameters
The nitrous oxide analogy method was used to estimate the CO2 solubility (Rinker and Sandall23) and the
CO2 diffusivity (Ashour et al.24) in the aqueous amine
solutions. The kinetics for CO2/MDEA were measured
by Rinker et al.,6 and the kinetics for the CO2/DEA
reaction were determined by Rinker et al.15 The diffusion coefficients of MDEA and DEA were estimated from
the diffusivity data of Hikita et al.25
We include here correlations that were used to
estimate various other reaction rate coefficients and
equilibrium constants that were obtained from the
literature. Values for the forward rate coefficient of
reactions 1 and 2, k1 and k2, respectively, were calculated from the following correlations, which were reported by Pinsent et al.26 for the temperature ranges of
273-311 K and 273-313 K, respectively:

The rates of absorption of CO2 into aqueous solutions


of DEA and MDEA were measured in a laminar-liquidjet absorber and a stirred-cell absorber. A schematic
drawing of the laminar-liquid-jet absorber is shown in
Figure 1. The laminar-liquid-jet absorber and its operation are described in detail by Rinker et al.15 The stirredcell absorber is shown schematically in Figure 2. The
absorption chamber is made of a 30.5 cm long, 10.1 cm
i.d. Pyrex cylinder and is enclosed in a constanttemperature heating jacket constructed from a 31 cm
long, 24 cm i.d. Lucite cylinder. The ends of the Lucite
cylinder are sealed with rubber O-rings between two
anodized aluminum flanges, and the glass cylinder is
sandwiched between two stainless steel flanges with the
ends sealed by Teflon gaskets. Cooling or heating water
is supplied to the jacket and recycled to a constanttemperature circulating bath. Two separate stainless
steel coils are placed in the heating/cooling jacket and
are used to control the temperatures of the liquid and
gas feeds. The liquid supply is introduced into the
chamber by a 0.635 cm o.d. stainless steel tube that can
slide in the vertical direction through the bottom flange
and can be locked in position by a Swagelok nut with
Teflon ferrules. The end of this tube is plugged, and the
liquid is discharged into the absorption chamber from
perforations on the side of the tube near the plugged
end. This assembly makes it possible to discharge the
liquid at different heights if so desired. The absorption
chamber is also equipped with four flat stainless steel
baffles, which help in reducing vortex formation and
promoting better mixing of the liquid phase. The baffles
are 12.7 cm long, 1.0 cm wide, and 0.1 cm thick and are
pinned to the bottom flange, and their top ends are
connected by a wire ring. The distance between the
baffles and the glass wall of the chamber is about 0.35
cm. When the liquid height in the chamber is 11.0 cm,
the baffles extend about 1.7 cm above the liquid surface.
The absorber has two concentric shafts that protrude
into the chamber through the top flange. The inner shaft
is 0.6 cm in diameter and 36.5 cm long and is made of
stainless steel. This shaft extends to a Teflon bushing
in the bottom flange and is supported at the top end by
a pin bearing held in a cup on a crossbar. There are

Ind. Eng. Chem. Res., Vol. 39, No. 11, 2000 4351

Figure 1. Schematic drawing of the laminar-liquid-jet absorber.


Table 1. Equilibrium Constant Correlations Used for Model Calculations
equilibrium
constant

temp range
(K)

equation

reference

293-573

Olofsson
and Hepler27

273-523

Read28

K14K16

log10(K16) ) 8909.483 - 142613.6/T


-4229195 log10(T) + 9.7384 T
-0.012 963 8T2 + (1.150 68 10-5)T3
-(4.602 10-9)T4
log10(K2K16) ) 179.648 + 0.019 244T
-67.341 log10(T) - 7 495.441/T
log10(K14K16) ) 6.498 - 0.023 8T - 2 902.4/T

273-323

K4K5K13K16
K13K16
K12K16

log10(K4K5K13K16) ) -10.549 2 + 1 526.27/T


log10(K13K16) ) -4.030 2 - 1 830.15/T + 0.0043T
log (K12K16) ) -14.01 + 0.018T

298-333
298-333
298-333

Danckwerts
and Sharma29
Barth et al.30
Barth et al.30
Barth et al.30

K16

K2K16

two liquid-phase impellers mounted on the shaft, one


with a diameter of 7.5 cm mounted at a height of 8.0
cm from the bottom flange, and another with a diameter
of 6.0 cm mounted at a height of 3.0 cm from the bottom
flange. The second shaft, which is a hollow stainless
steel tube 20.0 cm long with a 1.1 cm i.d., is supported
by a concentric tubular extension welded onto the top
flange and extends 10.0 cm into the chamber. Mounted
on the second shaft are two impellers for stirring the
gas phase, a 7.5 cm diameter impeller at 10.5 cm from
the top flange, and a 9.5 cm diameter impeller at 17.5
cm from the top flange. When the liquid height in the
chamber is 11.0 cm, the bottom edge of the second gas
impeller is 2.0 cm away from the gas-liquid interface.
Two sets of ball bearings, Teflon packings, and springloaded Teflon seals support the tubular shaft. Two
similar assemblies support the liquid shaft. The two
shafts are driven independently by two variable-speed
motors.
The liquid feed is pumped to a surge tank and then
through a calibrated rotameter and through a stainless
steel coil in the constant-temperature jacket. It is then

introduced into the absorption chamber through the


bottom flange and is discharged into the chamber by
seven perforations about 8 cm from the bottom of the
chamber. The liquid leaves the chamber through the
bottom flange and goes to a liquid-leveling reservoir.
The leveling device is similar to that used in the
laminar-liquid-jet absorber.
The gases pass through two sets of regulators (high
and low pressure) and then through mass flow controllers in order to maintain constant gas feed rates. The
gases subsequently pass through soap bubble meters so
that their volumetric flowrates can be measured. The
gases are next mixed in a T-fitting and fed to a
saturator. After the saturator, the overall gas flow rate
is measured with a bubble meter. The temperature of
the gas mixture is measured and recorded at this point
with a type-J thermocouple. The gas mixture then
passes through the stainless steel heating/cooling coil
in the constant-temperature jacket before being introduced into the absorption chamber through the tube in
the top flange. The gas mixture is sampled just before
it enters the absorption chamber and is analyzed with

4352

Ind. Eng. Chem. Res., Vol. 39, No. 11, 2000

Figure 2. Schematic drawing of stirred-cell absorber.

a gas chromatograph. The gas is discharged just above


the larger impeller through perforations near the end
of the feed tube. The gas exits the chamber through a
fitting in the top flange and is sampled for composition
by a gas chromatograph. The volumetric flow rate of the
exit gas is measured with a soap bubble meter and the
temperature of the exit gas is measured with a type-J
thermocouple.
The gas samples were analyzed using a model 5890
Hewlett-Packard gas chromatograph equipped with a
thermal conductivity detector. The column used in the
GC was a 6 ft. long, 0.085 in. i.d., stainless steel column
packed with Poropak Q with a mesh size of 80/100. The
run conditions were 70 C for 1.5 min, ramp to 90 C at
5 C per minute, cool, and equilibrate for 2 min.
Experimental Results
(A) Absorption Measurements at Short Contact
Times Using a Laminar-Liquid-Jet Absorber. According to the zwitterion mechanism, any base can
contribute to the deprotonation of the zwitterion. For
the aqueous DEA case, the potential bases are R2NH,
OH-, H2O, HCO, and CO. In an earlier study (Rinker
et al.15), the contributions of OH-, H2O, HCO, and CO
to the deprotonation of the zwitterion were found to be
insignificant for the reaction of CO2 with aqueous DEA
in the laminar-liquid-jet absorber. However, if CO2 is
absorbed into aqueous blends of DEA and MDEA, it is
possible that MDEA could contribute significantly to the

Table 2. CO2 Absorption Data in 10 wt % Blends of DEA


and MDEA Obtained in the Laminar-Liquid-Jet Absorber
at 25 C
[MDEA]
(kmol/m3)

[DEA]
(kmol/m3)

RCO2 105
measured
(kmol/m2s)

RCO2 105
predicted
(kmol/m2s)

0.844
0.808
0.692
0.586
0.532
0.232
0

0
0.040
0.173
0.293
0.355
0.696
0.959

1.22
1.29
1.45
1.65
1.71
2.40
3.28

1.22
1.35
1.44
1.65
1.76
2.69
3.32

deprotonation of the zwitterion. Littel et al.14 report that


MDEA makes a measurable contribution to the deprotonation of the zwitterion for CO2 absorption into
aqueous blends of DEA and MDEA.
To test the hypothesis that MDEA contributes to the
deprotonation of the zwitterion, the rates of CO2 absorption into aqueous blends of DEA and MDEA were
measured in the laminar-liquid-jet absorber at 25 C.
The total weight percents of the aqueous DEA/MDEA
blends were 10, 30, and 50 wt %, and the molar ratios
of DEA to MDEA were varied from zero to infinity. The
absorption data are listed in Tables 2-4. The rates of
CO2 absorption were predicted using the numerical
model developed in this work without taking into
account the contribution of MDEA to the deprotonation
of the zwitterion. The predicted and measured rates of
CO2 absorption are compared in the parity plot shown

Ind. Eng. Chem. Res., Vol. 39, No. 11, 2000 4353
Table 3. CO2 Absorption Data in 30 wt % Blends of DEA
and MDEA Obtained in the Laminar-Liquid-Jet Absorber
at 25 C
[MDEA]
(kmol/m3)

[DEA]
(kmol/m3)

RCO2 105
measured
(kmol/m2s)

RCO2 105
predicted
(kmol/m2s)

2.581
2.118
1.376
0.712
0
2.581
2.118
1.376
0.712
0

0
0.529
1.376
2.135
2.947
0
0.529
1.376
2.135
2.947

1.12
1.97
3.08
3.77
4.86
1.06
1.73
3.06
3.72
4.65

0.870
1.49
2.83
3.80
4.75
0.850
1.51
2.82
3.80
4.71

Table 4. CO2 Absorption Data in 50 wt % Blends of DEA


and MDEA Obtained in the Laminar-Liquid-Jet Absorber
at 25 C
[MDEA]
(kmol/m3)

[DEA]
(kmol/m3)

RCO2 105
measured
(kmol/m2s)

RCO2 105
predicted
(kmol/m2s)

4.217
4.217
3.722
3.719
3.722
3.034
3.034

0.202
0.202
0.772
0.771
0.772
1.517
1.517

0.753
0.868
1.10
1.15
1.09
1.67
1.50

0.692
0.783
1.11
1.12
1.09
1.72
1.74

in Figure 3. From this plot, it is clear that there is good


agreement between the measured rates of CO2 absorption and the rates of CO2 absorption predicted by the
model neglecting the contribution of MDEA to the
deprotonation of the zwitterion. The average deviation
of the predicted rates from the measured rates is 6.8%.
On the basis of these experiments, it appears that
MDEA does not significantly contribute to the deprotonation of the zwitterion.
A possible explanation for the difference in the results
of this study and the work of Littel et al.14 is that we
used a laminar-liquid-jet absorber whereas they used
a batch stirred-tank reactor. The laminar-jet absorber
operates at steady state, with a gas-liquid contact time
of about 0.005 s, whereas the reactor used by Littel et
al. operates under transient conditions and over much
longer gas-liquid contact times. For very short contact
times, the deprotonation reaction with MDEA may not
have time to contribute significantly to the overall rate
of absorption.
(B) Absorption Measurements at Long Contact
Times Using a Stirred-Cell Absorber. To predict
rates of CO2 absorption obtained using the stirred-cell
apparatus, accurate values of the liquid-phase mass
transfer coefficient for physical absorption must be
determined from experimental absorption data obtained
for a model system. The liquid-phase physical mass
transfer coefficient was determined by measuring the
concentration of dissolved CO2 in the liquid effluent
from the stirred-cell absorber using wet-chemical analysis. The liquids used for these experiments were pure
deionized water and pure poly(ethylene glycol) (PEG400).
The gas phase was pure CO2 saturated with the vapor
pressure of the liquid. The liquid feedstocks were
degassed at elevated temperatures. The concentration
of CO2 dissolved in the liquid was determined by
titration of samples of effluent. The water samples were
drawn slowly into a syringe and injected into an equal
volume of 0.05 M NaOH to convert all of the dissolved

Figure 3. Comparison of predicted and measured rates of CO2


absorption into aqueous blends of DEA and MDEA at 25 C in
the laminar-liquid-jet absorber. The total amine concentrations
were 10, 30, and 50 mass %.

CO2 into carbonate. The carbonate was precipitated


with barium chloride, and the remaining NaOH was
titrated with 0.05 M HCl. This method was reproducible
to within 1%, and it was accurate enough to measure
the small rates of CO2 absorption over liquid stirring
speeds of 30-140 rpm. For the pure PEG400 experiments, the samples were injected into twice their
volume of 0.05 M NaOH in order to convert the CO2 to
carbonate and dilute the PEG400 so that the samples
were mostly water (on a molar basis) for accurate pH
determination. Pure PEG400 was used to check the
viscosity dependence of the mass transfer coefficient as
it has a viscosity of 0.9527 P at 25 C and water has a
viscosity of 0.00895 P at 25 C.
The physical mass transfer coefficients were calculated from the measured data using the following
equations:

RA1 ) kl,1(C/A- CA0)


RA1 )

(58)

C0A L
As

(59)

where C/A is the interfacial concentration of CO2, C0A is


the bulk or effluent concentration of CO2, L is the
volumetric liquid flowrate, As is the area of the gasliquid interface (79.64 cm2), and k is the physical liquidphase mass transfer coefficient. The mass transfer
coefficients were correlated in dimensionless form as
follows:

Sh ) 0.0193Re0.845Sc0.5

(60)

where

Sh )

kl,1da
DA

Re )

FNLdi2

Sc )

FDA

(61)

The dimensionless mass transfer coefficients used to


obtain the correlation given above are plotted in
Figure 4.
(i) CO2 Absorption into Aqueous MDEA. Rates of CO2
absorption into aqueous solutions of MDEA were measured in the stirred-cell absorber at 25 C. The CO2
partial pressure was varied from 0.15 to 0.5 atm, and
the diluent was N2. The measured absorption rates are

4354

Ind. Eng. Chem. Res., Vol. 39, No. 11, 2000

Figure 4. Liquid-phase physical absorption mass transfer coefficient correlation.

Figure 5. Comparison of predicted and measured rates of CO2


absorption into aqueous MDEA at 25 C in the stirred-cell
absorber.

Figure 7. Comparison of predicted and measured rates of CO2


absorption into aqueous 50 mass % blends of DEA and MDEA at
25 C in the stirred-cell absorber.

with an average deviation of 10.2% from the measured


rates.
(iii) Absorption of CO2 into Aqueous Blends of DEA
and MDEA. Rates of absorption of CO2 into aqueous
blends of DEA and MDEA were measured in the stirredcell absorber at 25 C. The total amine concentration
was approximately 50 wt %, and the molar ratios of
DEA to MDEA were 0.050, 0.207, 0.250, and 0.500. The
model predictions are compared to the measured absorption rates in Figure 7. The average deviation of the
predictions from the measured rates of absorption is
17.6%. The predicted rates of absorption were made by
neglecting the contribution of MDEA to the deprotonation of the zwitterion. Some of the absorption rates are
overpredicted and some are underpredicted when the
MDEA contribution is neglected. As a result, the CO2
absorption data obtained in the stirred-cell absorber
support the findings for CO2 absorption into blends in
the laminar-liquid-jet absorber that MDEA does not
significantly contribute to the deprotonation of the
zwitterion.
Conclusions

Figure 6. Comparison of predicted and measured rates of CO2


absorption into aqueous DEA at 25 C in the stirred-cell absorber.

compared to the model predictions in the parity plot


shown in Figure 5. The average deviation of the model
predictions from the measured rates is 24%.
(ii) CO2 Absorption into Aqueous DEA. Rates of CO2
absorption into 10 wt % aqueous DEA were measured
in the stirred-cell absorber at 25 C. The CO2 partial
pressure was varied from 0.088 to 0.712 atm, and the
diluent was N2. The measured and predicted rates of
CO2 absorption are compared in the parity plot shown
in Figure 6. In this case, there is fairly good agreement
between the model predictions and the measurements,

The model developed in this work for the rates of


absorption of carbon dioxide into a aqueous mixed amine
solutions was found to agree reasonably well with the
experiments. The reaction between carbon dioxide and
the secondary amine, DEA, was described by the zwitterion mechanism in this model. For our experiments
with gas-liquid contact times varying from approximately 0.005 to 10 s, it appears that the tertiary amine,
MDEA, does not contribute significantly to the deprotation of the zwitterion. The only species that contributes significantly to the deprotonation of the zwitterion
is DEA, and thus, eqs 6-10 in the reaction model could
be deleted according to the results obtained in this
study.
The key physicochemical property needed for the
model calculations is the physical solubility of CO2. Any
uncertainty in this property translates to an equivalent
error in the predicted absorption rate.
Acknowledgment
This work was sponsored by the Gas Research Institute and the Gas Processors Association.

Ind. Eng. Chem. Res., Vol. 39, No. 11, 2000 4355

Notation
As ) gas-liquid interface area, m2
CA ) diameter of the laminar-liquid jet, m
Di ) coefficient of species i in the aqueous alkanolamine
solution, m2/s
DEA ) diethanolamine
DIPA ) diisopropanolamine
E1 ) factor of CO2, defined by eq 54
h0 ) spacing next to the gas-liquid interface for the
discretized spatial variable x, m
H1 ) physical equilibrium constant (Henrys law constant)
for CO2, H1 ) P/1/u/1, (kPa m3)/kmol
kapp ) rate coefficient for the reaction between CO2 and a
secondary or a primary alkanolamine, defined by eq 70,
m3/(kmol s)
kg,1 ) gas-phase mass transfer coefficient for CO2, kmol/
(kPa m2 s)
k0l,1 ) mass transfer coefficient for CO2 in the liquid phase,
defined by equation (52), m/s
ki ) rate coefficient of reaction i, s-1 for first-order
reactions, m3/(kmol s) for second-order reactions
k-i ) rate coefficient of reaction i
Ki ) constant for reaction i
l ) length of the laminar-liquid jet, m
L1 ) CO2 loading of the aqueous amine solution, (kmol
CO2)/(kmol amine)
MDEA ) N-methyldiethanolamine
MEA ) monoethanolamine
NL ) impeller speed, rev/s
P1 ) pressure of CO2 in the gas phase, kPa
P/1) partial pressure of CO2 in the gas phase, kPa
Q ) volumetric flow rate of the liquid, m3/s
Ri ) reaction rate of reaction i, kmol/(m3 s)
RA1 ) rate of absorption of CO2 per unit interfacial area,
defined by equation (53), kmol/(m2 s)
Re ) Reynolds number defined by eq 80
Sc ) Schmidt number defined by eq 80
Sh ) Sherwood number defined by eq 80
t ) independent time variable, s
T ) absolute temperature, K
ui ) concentration of species i in the liquid phase (which
is a function of x and t), kmol/m3
0
ui ) liquid bulk concentration of s pecies i (which is a
constant), kmol/m3
u/1 ) interfacial concentration of CO2 in the liquid, kmol/
m3
x ) independent spatial variable, m
Greek letters
) viscosity of the aqueous alkanolamine solution, kg/(m
s)
F ) density of the aqueous alkanolamine solution, kg/m3
) gas-liquid contact time, defined by eq 51 for a laminarliquid-jet absorber, s

Literature Cited
(1) Chakravarty, T.; Phukan, U. K.; Weiland, R. H. Reaction
of acid gases with mixtures of amines. Chem. Eng. Prog. 1985, 81
(4), 32-36.
(2) Donaldson, T. L.; Nguyen, Y. N. Carbon dioxide reaction
kinetics and transport in aqueous amine membranes. Ind. Eng.
Chem. Fundam. 1980, 19, 260-266.
(3) Haimour, N.; Bidarian, A.; Sandall, O. C. Kinetics of the
reaction between carbon dioxide and methyldiethanolamine.
Chem. Eng. Sci. 1987, 42, 1393-1398.
(4) Versteeg, G. F.; van Swaaij, W. P. M. On the kinetics
between CO2 and alkanolamines both in aqueous and nonaqueous
solutions II. Tertiary amines. Chem. Eng. Sci. 1988, 43, 587-591.

(5) Littel, R. J.; van Swaaij, W. P. M.; Versteeg, G. F. Kinetics


of carbon dioxide with tertiary amines in aqueous solution. AIChE
J. 1990, 36, 1633-1640.
(6) Rinker, E. B.; Ashour, S. S.; Sandall, O. C. Kinetics and
modelling of carbon dioxide absorption into aqueous solutions of
N-methydiethanolamine. Chem. Eng. Sci. 1995, 50, 755-768.
(7) Caplow, M. Kinetics of carbamate formation and breakdown.
J. Am. Chem. Soc. 1968, 90, 6795-6803.
(8) Danckwerts, P. V. The reaction of CO2 with ethanolamines.
Chem. Eng. Sci. 1979, 34, 443-446.
(9) Blauwhoff, P. M. M.; Versteeg, G. F.; van Swaaij, W. P. M.
A study on the reaction between CO2 and alkanolamines in
aqueous solutions. Chem. Eng. Sci. 1984, 39, 207-225.
(10) Versteeg, G. F.; van Swaaij, W. P. M. On the kinetics
between CO2 and alkanolamines both in aqueous and nonaqueous
solutions I. Primary and secondary amines. Chem. Eng. Sci. 1988,
43, 573-585.
(11) Versteeg, G. F.; Oyevaar, M. H. The reaction between CO2
and diethanolamine at 298 K. Chem. Eng. Sci. 1989, 44, 12641268.
(12) Versteeg, G. F.; Kuipers, J. A. M.; van Beckum, F. P. H.;
van Swaaij, W. P. M. Mass transfer with complex reversible
chemical reactions-II. Parallel reversible chemical reactions. Chem.
Eng. Sci. 1990, 45, 183-197.
(13) Glasscock, D. A.; Critchfield, J. E.; Rochelle, G. T. CO2
absorption/desorption in mixtures of methyldiethanolamine with
monoethanolamine or diethanolamine. Chem. Eng Sci. 1991, 46,
2829-2845.
(14) Littel, R. J.; Versteeg, G. F.; van Swaaij, W. P. M. Kinetics
of CO2 with primary and secondary amines in aqueous solutions
II. Influence of temperature on zwitterion formation and deprotonation rates. Chem. Eng. Sci. 1992, 47, 2037-2045.
(15) Rinker, E. B.; Ashour, S. S.; Sandall, O. C. Kinetics and
Modelling of Carbon Dioxide Absorption into Aqueous Solutions
of Diethanolamine. Ind. Eng. Chem. Res. 1996, 35, 1107-1114.
(16) Sada, E.; Kumazawa, H.; Han, Q.; Matsuyama, H. Chemical kinetics of the reaction of carbon dioxide with ethanolamines
in nonaqueous solvents. AIChE J. 1985, 31, 1297-1303.
(17) Bosch, H.; Versteeg, G. F.; van Swaaij, W. P. M. Kinetics
of the reaction of CO2 with the sterically hindered amine 2-aminomethylpropanol at 298 K. Chem. Eng. Sci. 1990, 45, 1167-1173.
(18) Alper, E. Reaction mechanism and kinetics of aqueous
solutions of 2-amino-2-methyl-1-propanol and carbon dioxide. Ind.
Eng. Chem. Res. 1990, 29, 1725-1728.
(19) Hanna, O. T.; Sandall, O. C. Computational Methods in
Chemical Engineering; Prentice Hall: Englewood Cliff, NJ, 1994.
(20) Higbie, R. The rate of absorption of a pure gas into a still
liquid during short periods of exposure. Trans. Am. Inst. Chem.
Eng. 1935, 31, 365.
(21) Danckwerts, P. V. Gas-Liquid Reactions; McGraw-Hill:
New York, 1970.
(22) Petzold, L. R. A Description of DASSL: A differential/
algebraic system solver. In Scientific Computing; Stepleman, R.
S., Carver, M., Peskin, R., Ames, W. F., Vichnevetsky, R., Eds.;
North-Holland; Amsterdam, 1983; pp 65-68.
(23) Rinker, E. B.; Sandall, O. C. Solubility of nitrous oxide in
aqueous solutions of methyldiethanolamine, diethanolamine and
mixtures of methyldiethanolamine and diethanolamine. Chem.
Eng. Commun. 1996, 144, 85-94.
(24) Ashour, S. S.; Rinker, E. B.; Sandall, O. C. Correlations
for estimating the diffusivities of nitrous oxide in aqueous solutions
of diethanolamine and n-methyldiethanolamine and the solution
densities and viscosities. Chem. Eng. Commun. 1997, 161, 15-24.
(25) Hikita, H.; Ishikawa, H.; Uku, K.; Murakami, T. Diffusivities of mono-, di-, and tri-ethanolamines in aqueous solutions.
J. Chem. Eng. Data 1980, 25, 324-325.
(26) Pinsent, B. R. W.; Pearson, L.; Roughton, F. J. W. The
kinetics of combination of carbon dioxide with hydroxide ions.
Trans. Faraday Soc. 1956, 52, 1512-1520.
(27) Oloffson, G.; Hepler, L. G. Thermodynamics of ionization
of water over wide ranges of temperature and pressure. J. Solution
Chem. 1975, 4, 127.
(28) Read, A. J. The first ionization constant of carbonic acid
from 25 to 250 and to 2000 bar. J. Solution Chem. 1975, 4, 5370.
(29) Danckwerts, P. V.; Sharma, M. M. The absorption of carbon
dioxide into solutions of akalis and amines (with some notes on
hydrogen sulphide and carbonyl sulphide). Chem. Eng. 1966,
October, 44, CE244-CE280.

4356

Ind. Eng. Chem. Res., Vol. 39, No. 11, 2000

(30) Barth, D.; Tondre, C.; Lappai, G.; Delpuech, J.-J. Kinetic
study of carbon dioxide reaction with tertiary amines in aqueous
solutions. J. Phys. Chem. 1981, 85, 3660-3667.
(31) Cotton, F. A.; Wilkinson, G. Advanced Inorganic Chemistry;
Interscience Publishers: New York, 1966.
(32) Smith, J. M.; Van Ness, H. C. Introduction to Chemical
Engineering Thermodynamics, 4th ed.; McGraw-Hill: New York,
1987.

(33) Lide, D. R., Ed. CRC Handbook of Chemistry and Physics;


71st ed.; CRC Press: Boca Raton, FL, 1990.

Received for review November 22, 1999


Revised manuscript received July 17, 2000
Accepted July 20, 2000
IE990850R

You might also like