You are on page 1of 24

This article was downloaded by: [University of Arizona]

On: 27 November 2012, At: 05:26


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Chemical Engineering Communications


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/gcec20

GAS–LIQUID REACTION KINETICS: A


REVIEW OF DETERMINATION METHODS
a a
Prakash D. Vaidya & Eugeny Y. Kenig
a
Department of Biochemical and Chemical Engineering, University
of Dortmund, Dortmund, Germany
Version of record first published: 03 Aug 2007.

To cite this article: Prakash D. Vaidya & Eugeny Y. Kenig (2007): GAS–LIQUID REACTION KINETICS: A
REVIEW OF DETERMINATION METHODS, Chemical Engineering Communications, 194:12, 1543-1565

To link to this article: http://dx.doi.org/10.1080/00986440701518314

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation
that the contents will be complete or accurate or up to date. The accuracy of any
instructions, formulae, and drug doses should be independently verified with primary
sources. The publisher shall not be liable for any loss, actions, claims, proceedings,
demand, or costs or damages whatsoever or howsoever caused arising directly or
indirectly in connection with or arising out of the use of this material.
Chem. Eng. Comm., 194:1543–1565, 2007
Copyright # Taylor & Francis Group, LLC
ISSN: 0098-6445 print/1563-5201 online
DOI: 10.1080/00986440701518314

Gas–Liquid Reaction Kinetics: A Review


of Determination Methods

PRAKASH D. VAIDYA AND EUGENY Y. KENIG


Department of Biochemical and Chemical Engineering, University
of Dortmund, Dortmund, Germany
Downloaded by [University of Arizona] at 05:26 27 November 2012

The aim of this article is to provide comprehensive insight into the determination and
interpretation of reaction kinetics of two-phase (gas–liquid) systems. Various
aspects of the methodologies used for the measurements of kinetic parameters (such
as equipment design, corresponding theoretical background, main steps, advantages,
and limitations) are discussed in detail. In addition, an illustrating example is
provided based on an industrially relevant absorption system.

Keywords Absorption; Gas-liquid reactions; Reaction kinetics

Introduction
Gas-liquid reactions have received considerable attention by both academicians and
industry. Some examples of industrial relevance are absorption of CO2 into aqueous
alkanolamines, absorption of H2S into ferric sulphate solutions, absorption of NO
or NO2 into solutions containing reactive species, absorption of O2 in sodium sul-
phite solutions, hydrogenation of olefins using homogeneous catalysts, recovery of
iso-butylene from C4 streams (by selective absorption in sulphuric acid), and chlori-
nation in aqueous solutions containing phenol and substituted phenols. A detailed
knowledge of reaction kinetics for such reaction systems is essential for the design
of suitable reactors and prediction of reactor performance. Absorption experiments
are therefore performed on a laboratory scale to facilitate fundamental understand-
ing of these reactions, to study the effect of the chemical reaction on mass transfer
rates, and to describe the reaction rate as a function of process variables, e.g., tem-
perature and reactant concentration. Such bench-scale kinetic experiments are
performed in the range of temperatures and concentrations typical for industrial
absorbers. Danckwerts (1970) and Doraiswamy and Sharma (1984) have earlier out-
lined the theory of mass transfer with chemical reaction and suggested models that
can be employed to study the kinetics of gas–liquid reaction systems.
Experimental techniques used to obtain the kinetic data for such systems may
differ. Further, experimental measurements may at times be erroneous. In addition,
different theoretical models can be used for the parameter treatment. Hence, even for
identical conditions, some deviations in estimated kinetic parameters may arise in
different investigations. Aboudheir et al. (2003) compiled literature data on the reac-
tion between CO2 and aqueous monoethanolamine (MEA) and showed that the

Address correspondence to Eugeny Y. Kenig, Department of Biochemical and Chemical


Engineering, University of Dortmund, 44227 Dortmund, Germany. E-mail: e.kenig@bci.
uni-dortmund.de

1543
1544 P. K. Vaidya and E. Y. Kenig

reaction rate constant reported by different techniques at 298 K varied from 3880
to 8400 m3=(kmol s). They attributed this wide variation in the rate constant to
uncertainties in values of the physical properties used, as well as to the existence
of interfacial turbulence in some types of absorbers, lack of knowledge of the exact
gas–liquid interfacial area, and the assumption of a pseudo-first-order reaction.
It would therefore be desirable to have a look at existing methods used in labora-
tories to study gas–liquid reaction kinetics. Two types of contactors have primarily
been used. In the first category, the flow pattern is well defined, for instance, in
wetted-wall columns, wetted-sphere contactors, and laminar jet absorbers. In the
second category, the flow pattern is not well established. Examples are stirred cells,
stirred contactors, and mechanically agitated contactors. Except for the mechani-
cally agitated contactor, the mass transfer interfacial area for all of the above devices
is geometrically simple and hence known. In a laminar jet absorber or in a stirred cell
Downloaded by [University of Arizona] at 05:26 27 November 2012

reactor, gaseous components are transported into the liquid phase and hence equi-
librium and mass transfer considerations should be taken into account. However,
in experiments without phase transition, such as in the rapid mixing method or
stopped-flow method, the gas is dissolved physically in a liquid and then mixed with
the liquid absorbent. Mass transfer is not considered in such methods, and the kin-
etic parameters are determined from the rate of the homogeneous reaction between a
solute and a reactant.
In the present article, we give an overview of theory and experimental methods
for studying reaction kinetics. Furthermore, an illustration of the application of
different methods is given for an industrially relevant absorption system: CO2
absorption by aqueous diethanolamine (DEA) solutions.

Theory

Mass Transfer Model


The two-film model, originally proposed by Lewis and Whitman (1924), is a simpli-
fied model used to describe mass transfer at the gas–liquid interface (Figure 1).

Figure 1. Schematic of the two-film model adapted from Kenig et al. (2003).
Gas–Liquid Reaction Kinetics 1545

It is assumed that the gas and liquid are in equilibrium at the interface and that thin
films separate the interface from the bulk of both contacting phases. Further, trans-
fer occurs within these films by molecular diffusion alone. Outside the films, in the
bulk fluid phases, the level of turbulence is so high that there is no composition
gradient at all. The liquid-side mass transfer coefficient kL represents the amount
of solute (here denoted by A) transferred through the liquid film per unit time, unit
area, and driving force unit in terms of liquid concentration. Similarly, the gas-side
mass transfer coefficient kG is the amount of solute transferred through the gas film
per unit time, unit area, and driving force unit in terms of pressure. Due to the one-
dimensional nature of the film model, the amount of solute transferred from the bulk
gas phase to the interface must equal the amount transferred from the interface to
the bulk liquid, and hence, the following relation holds:
 
Downloaded by [University of Arizona] at 05:26 27 November 2012

RA ¼ kG ðpA  pAI Þ ¼ kL CIA  CA ð1Þ

Effect of Chemical Reaction in the Liquid Phase


In the case of very slow reactions between a solute gas and a reactant in the liquid
phase, the dissolved solute molecules are transported deep into the bulk of the liquid
before the reaction can start. Hence the overall absorption rate is not increased sig-
nificantly due to chemical reaction. The rate of mass transfer is governed by the con-
centration difference between the interface and the bulk liquid. In the case of very
fast reactions, the solute molecules only slightly penetrate into the liquid phase
before the reaction begins. The diffusion path of the solute is extremely small when
compared to the distance it travels in simple physical absorption.
The reaction velocities may vary widely, from very low (very slow reactions) to
infinitely high (instantaneous reactions). It is possible to differentiate between very
slow, slow, fast, and instantaneous reactions, with respect to physical mass trans-
port, depending on the value of the Hatta number (Doraiswamy and Sharma,
1984). The latter represents the ratio of maximal possible reaction and mass transfer
rates. For an irreversible reaction of mth order with respect to the solute A and nth
order with respect to the reactant B in the liquid phase, the Ha number is expressed
as (Hikita and Asai, 1964):
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 I m1
2
mþ1 DA kmn CA ðCB Þn
Ha ¼ ð2Þ
kL

The most frequently used way to describe the enhancing effect of chemical
reactions on mass transport is to apply the so-called enhancement factor E. A
number of different methods for the description of the enhancement factor can
be found in the literature (van Swaaij and Versteeg, 1992). Depending upon the
mass transfer model (the two-film model (Lewis and Whitman, 1924), the pen-
etration model (Higbie, 1935), and the surface renewal theory (Danckwerts,
1970)), a variety of enhancement factor expressions have been developed. For
the case when the reaction of the solute gas occurs entirely in the liquid film
(Figure 2; see Doraiswamy and Sharma, 1984), the enhancement factor is equal
to the Ha number. The reactant B is present in excess and thus its concentration
1546 P. K. Vaidya and E. Y. Kenig
Downloaded by [University of Arizona] at 05:26 27 November 2012

Figure 2. Concentration profile when reaction occurs entirely in the liquid film, according to
Doraiswamy and Sharma (1984).

at the gas–liquid interface is almost the same as that in the bulk liquid. If the reac-
tion is represented by
A þ zB ! Products ð3Þ

and the reaction kinetics is described by


 m
r ¼ kmn CIA ðCB Þn ð4Þ

the specific rate of absorption can be expressed as (Doraiswamy and Sharma,


1984):
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2  mþ1
RA ¼ kL CIA E ¼ DA kmn CIA ðCB Þn ð5Þ
mþ1

Equation (5) is valid for 3 < Ha < (Ei  1) where Ei is the enhancement factor for
an instantaneous chemical reaction (Danckwerts, 1970),
!
C B DB
Ei ¼ 1 þ ð6Þ
zCIA DA

Estimation of Reaction Kinetics


Equation (5) can be used to estimate the reaction kinetic parameters, that is, the
reaction rate constant and the reaction order with respect to the reactants. If absorp-
tion rates at different values of CIA are measured (while keeping CB constant), the
value of m can be estimated from the slope of the straight line fitting the data of
log RA versus log CIA . The concentration of the dissolved gas in the liquid at the
interface can be calculated using Henry’s coefficient, HA:
CIA ¼ HA pAI ð7Þ
Gas–Liquid Reaction Kinetics 1547

1=2
If the variation in RA with CB is studied, another plot, log fRA =½ðCIA Þmþ1=2 DA g
versus log CB, enables the estimation of the value of n. The value of kmn can then
1=2
be found from a plot of log RA versus log f2=m þ 1 ðDA ÞðCIA Þðmþ1Þ=2 ðCB Þn=2 g.
Shen et al. (1991) used such plots to investigate the kinetics of the reaction between
CO2 and 2-piperidineethanol (PE) in aqueous solutions.
For the case when m ¼ n ¼ 1, the slope of the plot of ½RA =ðCIA Þ2  versus CB
yields the reaction rate constant. It follows from Equation (5) that prior knowledge
of physical parameters, such as solubility and diffusivity of the solute in the liquid
phase, is essential to determine kmn. Further, the specific rate of absorption is inde-
pendent of kL when the reaction occurs entirely in the film. It is worth noting that
uncertainties in absorption rate measurements, or in solubility and diffusivity
measurements, will lead to errors in estimation of the reaction rate constant.
Downloaded by [University of Arizona] at 05:26 27 November 2012

Experimental Methods for Determination of Gas-Liquid Reaction Kinetics


In this section, various aspects of the equipment used for the measurements of
kinetic parameters are discussed in detail. It should be noted that the wetted-wall
column, the wetted-sphere absorber, and the laminar jet apparatus can also be used
to estimate the solute diffusivity in the liquid phase.

Stirred Cell Reactor


A schematic of this reactor type operated in the batch mode is depicted in Figure 3.
The reactor consists of three parts: the lower flange on the liquid side, the upper gas-
side flange, and a jacketed glass tube between them. A pressure transducer mounted
on the gas-side flange, coupled with a data acquisition system, enables measurement
of the total pressure inside the reactor. Two temperature sensors are used to measure
the temperature of the gas and liquid phase. The reactor is equipped with inlet and
outlet ports for the gas and liquid. Both the gas and liquid phases are stirred by two
impellers, which are mounted on the same shaft. Baffles present inside the reactor
prevent vortex formation in the liquid. Water at the desired temperature is circulated
through the reactor jacket to ensure isothermal conditions.
Based on the conservation principle and using the ideal gas law, the following
equations can be derived for absorption of solute gas in a stirred cell:
dnA;G VG dpA
¼ 0 ð8Þ
dt R T dt
dnA;L dnA;G
¼ ð9Þ
dt dt
The volumetric rate of absorption of the solute gas is given by the following
correlation:
1 dnA;L VG dpA
RA a ¼ ¼ ð10Þ
VL dt VL R0 T dt
A series of experiments should be conducted at different temperatures and concen-
trations to determine reaction kinetics. In each experiment, a certain amount of sol-
ute gas should be rapidly introduced into the reactor, and the pressure decrease
caused by reaction should be recorded. From this information, the absorption rate
1548 P. K. Vaidya and E. Y. Kenig
Downloaded by [University of Arizona] at 05:26 27 November 2012

Figure 3. Schematic of the stirred cell reactor.

at a particular temperature and liquid-phase composition can be determined. The


solute gas concentration can be varied by changing the total pressure of the system.
Alternatively, an inert gas can also be used to change the partial pressure and hence
the concentration of the solute gas. However, the gas-side resistance may be signifi-
cant when the solute is mixed with an inert gas. Kucka et al. (2003) compared differ-
ent procedures to obtain the relevant rate parameters from the experimental data
obtained, using a stirred cell reactor. Some recent kinetic studies carried out using
a stirred cell reactor are listed in Table I (Kucka et al., 2002, 2003; Ebrahimi et al.,
2003; Long et al., 2005; Camacho et al., 2005; Vaidya and Mahajani, 2005).
The main advantage of the stirred cell is that the rates of absorption can be
measured using a liquid with a single, known composition. The analysis of the
liquid phase is not required if the reaction mechanism is known, because the pres-
sure decrease is the only decisive factor for estimation of kinetic parameters. Thus,
this technique is ideally suited for toxic liquids. This contactor can also be used to
estimate the solubility of the solute in the liquid phase. A disadvantage of this
method is that the stirrer must be driven by a constant speed motor. In addition,
the value of kL might be sensitive to the depth of immersion of the stirrer blades in
the liquid.
A similar type of contactor is a Lewis cell. Cadours et al. (2007) studied the acid
gas-aqueous DEA reaction system in such a reactor. The total pressure inside the
reactor was controlled by nitrogen injection in order to control the gas-side mass
transfer resistance. Amararene and Bouallou (2004) also studied kinetics of the
Downloaded by [University of Arizona] at 05:26 27 November 2012

Table I. Some recent kinetic studies carried out in a stirred cell reactor
Solute gas Absorbent Temp, C CB, kmol=m3 m n kmn Ref.

CO2 Aqueous NaOH, 20–50 0.075–0.85 1 1 Kucka et al. (2002)


aqueous KOH
CO2 Aqueous MEA 20–51 20 wt% 1 1 Kucka et al. (2003)
H 2S Aqueous ferric 25–65 0.025–0.8 1 1 exp[9.38–(2636.6=T)] Ebrahimi et al. (2003)

1549
sulphate m3=(kmol s)
NO Aqueous cobalt 35–65 >0.02 Long et al. (2005)
ethylene diamine
CO2 Aqueous DIPA 15–40 0.0413–1.9 1 2 1.41–11.8 m6=(kmol2 s) Camacho et al. (2005)
CO2 Aqueous MEA þ 30 1.5–2.5 1 1 11203 m3=(kmol s) Vaidya and Mahajani
NMP þ DEG (2005)
1550 P. K. Vaidya and E. Y. Kenig

reaction of carbonyl sulphide with aqueous solutions containing DEA and methyl-
diethanolamine (MDEA) in a Lewis cell.

Wetted-Wall Column
A schematic of this apparatus is depicted in Figure 4 (see Cullinane and Rochelle,
2006). It has the form of a vertical cylindrical stainless steel tube; the liquid is
pumped through the interior of the tube. The liquid overflows and is evenly distrib-
uted as a thin film along the outer surface of the tube. Gas enters near the base of the
column and countercurrently contacts the liquid. The absorption chamber is a glass
cylinder, enclosed by a jacket. Water is circulated through the jacket to maintain iso-
thermal conditions. The solute gas is mixed with an inert gas to give a large range of
partial pressures. The absorption rate is measured by the product of the gas flow rate
Downloaded by [University of Arizona] at 05:26 27 November 2012

and the difference between the inlet and outlet solute gas-phase concentrations.
Many researchers have studied reaction kinetics using such type of contactor
(Yih and Shen, 1988; Xiao et al., 2000; Horng and Li, 2002; Liao and Li, 2002;
Yoon et al., 2002, 2003; Sun et al., 2005; Cullinane and Rochelle, 2006; Al-Juaied
and Rochelle, 2006; Lee et al., 2007). These studies are summarized in Table II. A
disadvantage of this method is a possible appearance of ripples in the film, which
may lead to erroneous results. Rippling on the surface of the liquid can, however,
be avoided by the addition of surface-active agents (Danckwerts, 1970). Another
drawback, known as the stagnant layer end effect, is the formation of a rigid
inactive film of surface-active material that builds up at the bottom of the column.
The total film height therefore needs to be corrected for this end effect (Yoon et al.,
2002).

Figure 4. Schematic of the wetted-wall column according to Cullinane and Rochelle (2006).
Downloaded by [University of Arizona] at 05:26 27 November 2012

Table II. Some recent kinetic studies carried out in a wetted-wall column
Solute gas Absorbent Temp, C CB, kmol=m3 m n kmn Ref.

CO2 Aqueous solutions 40 0.258–3 1 1 1270 m3=(kmol s) Yih and Shen (1988)
of AMP
CO2 Aqueous solutions 30–40 AMP (1.5–1.7) þ 1 1 1258.4–7621.9 1=s Xiao et al. (2000)
of AMP þ MEA MEA (0.1–0.4)
CO2 Aqueous solutions 30 MEA (0.1) þ TEA (0.5) 1 1 190.3 1=s Horng and Li (2002)
of MEA þ TEA
CO2 Aqueous solutions 30 MEA (0.1) þ 1 1 118.7 1=s Liao and Li (2002)
of MEA þ MDEA MDEA (1.0)

1551
CO2 Aqueous AEPD 32 5–25 wt% 1 1 378 m3=(kmol s) Yoon et al. (2002)
CO2 Aqueous AMPD 30 5–25 wt% 1 1 382 m3=(kmol s) Yoon et al. (2003)
CO2 Aqueous solutions 30 AMP–1.5 þ PIP–0.1 1 1 4473 1=s Sun et al. (2005)
of AMP þ PIP
CO2 Aqueous solutions 25–110 K2CO3 0–3.1 M Cullinane and Rochelle
of K2CO3 þ PIP PIP 0.45–3.6 M (2006)
CO2 Aqueous mixtures 15–40 11 wt% MOR þ 1 2 Al-Juaied and Rochelle
of DGA þ MOR 53 wt% DGA (2006)
CO2 Aqueous sodium 30–50 1–3.5 1 1 218–1034 m3=(kmol s) Lee et al. (2007)
glycinate
1552 P. K. Vaidya and E. Y. Kenig

If physical absorption of a sparingly soluble gas in an initial gas-free liquid


occurs for a short contact time, the absorption rate by the wetted surface can be
described by the Higbie penetration theory (Danckwerts, 1970):
sffiffiffiffiffiffiffi
DA
RA ¼ 2 CIA ð11Þ
ptc

where the contact time tc is related to the wetted-wall column hydrodynamics as


  2=3  1=3
2h pd 3g
tc ¼ ð12Þ
3 L qg

Equation (11) can be used to estimate solute diffusivity in the liquid phase.
Downloaded by [University of Arizona] at 05:26 27 November 2012

Wetted-Sphere Absorber
This type of contactor is presented in Figure 5 (see Seo and Hong, 2000). The
absorption chamber is a jacketed glass cylinder, placed between two steel flanges.
The absorbent liquid flows over the surface of a sphere. The sphere is supported
by a vertical tube that passes through its center. The liquid enters through the tube
and emerges at the top of the sphere. It runs down the sphere surface and outer side,
making a smooth film. It flows down a short length of the tube and is collected in a
liquid receiver. The liquid flow rate is varied to make the film stable at the surface of
the sphere. The length between the bottom of the sphere and the liquid height in the
receiver is controlled by a liquid level controller. The absorption rates can be esti-
mated by using a soap-film meter. The advantages of this method as compared to
the wetted-wall column are that the stagnant layer end effects as well as surface rip-
pling are reduced. Many researchers have studied reaction kinetics using this type of
absorber (see Table III) (Savage et al., 1980; Tseng et al., 1988; Tomcej and Otto,
1989; Al-Ghawas et al., 1989b; Davis and Sandall, 1993; Rinker et al., 1995;
Dehouche and Lieto, 1995; Seo and Hong, 2000).

Figure 5. Wetted-sphere absorber according to Seo and Hong (2000).


Downloaded by [University of Arizona] at 05:26 27 November 2012

Table III. Some recent kinetic studies carried out in a wetted-sphere absorber
Solute gas Absorbent Temp, C CB, kmol=m3 m n kmn Ref.

CO2 K2CO3 solutions 40–110 1 1 Savage et al. (1980)


CO2 DEA promoted 50–103 25% K2CO3 þ 1 Tseng et al. (1988)
carbonate solutions DEA (2–5%)
CO2 Aqueous MDEA 25.1–74.9 20–40% MDEA 1 1 1.615  108 Tomcej and Otto (1989)
exp(5134=T)
COS Aqueous MDEA 20–40 1.259–2.599 1 1 4198.74 Al-Ghawas et al. (1989b)

1553
exp(4575.80=T)
CO2 DEA þ DIPA in PEG 20–40 1 Davis and Sandall (1993)
CO2 Aqueous MDEA 20–69 10–30% MDEA 1 1 2.91  107 Rinker et al. (1995)
exp(4579=T)
CO2 Aqueous Na2CO3 24 0.6 M 1 1 0.0195 m3=(mol s) Dehouche and Lieto
O2 Aqueous Na2SO3 25 0.189 M 2 1 0.073 (m3=mol)2=s (1995)
CO2 Aqueous AMP þ PIP 30, 40 AMP 0.55–3.35 1 1 Seo and Hong (2000)
PIP 0.058–0.233
1554 P. K. Vaidya and E. Y. Kenig

Davidson and Cullen (1957) presented a solution for the problem of physical gas
absorption by a laminar liquid film flowing over a sphere. For a small depth of pen-
etration, the rate of absorption can be predicted by a series expansion of the form:
h X i
RA ¼ LðCIA  CA Þ 1  bi expðci aÞ ð13Þ

where
 
2pg 1=3 7=3 4=3
a ¼ 3:36p rs L DA ð14Þ
3n
When absorption of a gas into a spherical liquid film is accompanied by a reac-
tion that occurs under the pseudo–first-order fast reaction conditions (m ¼ n ¼ 1
and 3 < Ha < (Ei  1)), the rate of absorption is given by (Tomcej and Otto, 1989):
Downloaded by [University of Arizona] at 05:26 27 November 2012

pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
RA ¼ 4pr2s DA kov ðCIA  CA Þ ð15Þ

Lynn et al. (1955) studied the hydrodynamics of laminar liquid films flowing over
a sphere and showed that the film thickness at the equator of the spherical film is
given by:
 
3nL 1=3
D¼ ð16Þ
2prs g
The gas–liquid contact time is given by (Seo and Hong, 2000):
 
2 3n 1=3 2=3
tc ¼ 3:441prs L ð17Þ
2prs g

Hemispherical Contactor
Jamal et al. (2006) studied the kinetics of CO2 absorption and desorption in aqueous
alkanolamine solutions using the hemispherical type of contactor. In principle,
it is similar to the wetted-sphere absorber. However, the full sphere is replaced by
a hemisphere, i.e., the liquid descends as a film over a hemispherical surface. Such
a contactor reduces the surface rippling that is encountered in the lower half of
the full spherical unit. Jamal (2002) showed that the rate of physical absorption over
a hemispherical surface is given by:
 2 1=2
Lrs DA
RA ¼ 3:1774 ðCIA  CA Þ ð18Þ
D
and the contact time is given by:
 
D
tc ¼ 1:5848 pr2s ð19Þ
L
Similar to Equations (11) and (13), Equation (18) can also be used to calculate solute
diffusivity in the liquid phase from the measured absorption rates.
Gas–Liquid Reaction Kinetics 1555

Laminar Jet Absorber


Aboudheir et al. (2003) studied the kinetics of the reaction between CO2 and high
CO2-loaded concentrated aqueous MEA solutions in a laminar jet absorber. A
scheme of the device, which was used in this study, is shown in Figure 6. The
absorption chamber is a glass cylinder, enclosed by a jacket. Liquid is introduced
into this chamber through a delivery tube. A nozzle is mounted at the end of this
tube. The nozzle has the form of a circular hole in a steel sheet. A liquid jet of a
very small diameter comes from this nozzle, flows downward through an atmos-
phere of the absorbed gas and is collected in a receiver. The receiver is a capillary
hole drilled in a rod. The liquid height in the receiver is controlled by a liquid level
controller. A soap-film meter is used to measure the absorption rates. Some recent
kinetic studies carried out using such a device are listed in Table IV (Rinker et al.,
1996, 2000; Aboudheir et al., 2003, 2004; Ramachandran et al., 2006). A disadvan-
Downloaded by [University of Arizona] at 05:26 27 November 2012

tage of this method is that the jet length cannot be varied over a wide range,
thereby restricting the range of gas–liquid interfacial area (Astarita et al., 1983;
Aboudheir et al., 2004).
The solute gas concentration should be changed by changing the total pressure
in the apparatus. It is not desirable to introduce an inert gas to change partial pres-
sure of the solute gas, since this may introduce uncertainties regarding the gas-side
resistance (Doraiswamy and Sharma, 1984). The contact time is determined by the
jet length and diameter and the liquid flow rate and is independent of the density
and viscosity of the liquid. It is given by the following relation (Rinker et al., 2000):
pd2j hj
tc ¼ ð20Þ
4L

Figure 6. Schematic of the laminar jet absorber according to Aboudheir et al. (2003).
Downloaded by [University of Arizona] at 05:26 27 November 2012

Table IV. Some recent kinetic studies carried out in a laminar jet absorber
Solute gas Absorbent Temp, C CB, kmol=m3 m n kmn Ref.

CO2 Aqueous DEA 20–70 0.25–2.8 1 1 Rinker et al. (1996)


CO2 Aqueous mixtures of DEA 25 MDEA (0–4.2) þ Rinker et al. (2000)
and MDEA DEA (0–2.9)

1556
CO2 Aqueous MEA 20–60 3–9 Aboudheir et al. (2003)
CO2 Aqueous MEA 20–40 0.193–3.158 1 1 4615–10119 Aboudheir et al. (2004)
m3=(kmol s)
CO2 Aqueous solutions of MEA 25–60 MDEA (23–27%) þ 1 1 Ramachandran et al.
and MDEA MEA (3–7%) (2006)
Gas–Liquid Reaction Kinetics 1557

According to the penetration theory, the average rate of absorption (kmol=s) is


given by:
RA ¼ 4CIA ðDA hj LÞ1=2 ð21Þ
A series of experiments at various flow rates and jet lengths can be performed. A
plot of RA versus ðhj LÞ1=2 is a straight line passing through the origin, having a slope
1=2
equal to 4CIA DA . Al-Ghawas et al. (1989a) measured the diffusivity of N2O in
aqueous MDEA solutions using this method.

Disk Column
A disk column is depicted in Figure 7. Xu et al. (1992) investigated the kinetics of
the reaction of CO2 with aqueous MDEA solutions containing piperazine (PIP) in
Downloaded by [University of Arizona] at 05:26 27 November 2012

this contactor, which was originally developed by Stephens and Morris (1951). The
column is a glass tube, enclosing a number of circular disks, which are held together
by a vertical wire. The solute gas enters at the bottom and flows upward through this
column. Liquid is introduced from the top. After direct contact with the gas, the liquid
is collected in a receiver and recycled to the column in order to increase the solute
content in the solution. Water is circulated to maintain isothermal conditions in the
column. The absorption rates at various time intervals can be calculated by using a
soap-film meter. Liquid samples can be collected at these time intervals and the solute
content in the absorbent can be analyzed. Knowledge of the disk diameter, disk thick-
ness, and number of disks is essential. A disadvantage of this method is that the com-
position of the gas and liquid may vary appreciably over the length of the column,
thereby making a direct measurement of absorption rates with known liquid and
gas compositions impossible (Danckwerts, 1970). A few kinetic studies carried out
in such a device are listed in Table V (Xu et al., 1992; Zhang et al., 2001, 2002).

Figure 7. Schematic of the disk column according to Xu et al. (1992).


Downloaded by [University of Arizona] at 05:26 27 November 2012

Table V. Some recent kinetic studies carried out in a disk column


Solute gas Absorbent Temp, C CB, kmol=m3 m n kmn Ref.

CO2 Aqueous MDEA þ PIP 30–70 MDEA (1.75–4.21) þ PIP 1 1 Xu et al. (1992)
(0.041–0.21)

1558
CO2 Aqueous MDEA þ PIP 30–70 MDEA (2.35–2.77) þ PIP 1 1 kPIP ¼ 4  1010 Zhang et al. (2001)
(0.23–0.65) exp(4059.4=T)
CO2 Aqueous MDEA þ DEA 40–70 MDEA (2.445–2.695) þ DEA 1 1 ln kDEA ¼ Zhang et al. (2002)
(0.191–0.555) 24.515–(5411.3=T)
Gas–Liquid Reaction Kinetics 1559

Homogeneous Experimentation Techniques


In order to avoid uncertainties concerning the physical properties such as diffusivity
and gas solubility in the liquid phase, homogeneous experimentation techniques can
be used for studying reaction kinetics. Hikita et al. (1977) investigated the kinetics
of the reactions of CO2 with aqueous alkanolamines by using one such technique,
namely, the rapid mixing method, which was developed by Hartridge and Roughton
(1923). Aqueous solutions of CO2 and alkanolamines were mixed together in a reactor,
and the intrinsic rate of the homogeneous reaction was calculated from the tempera-
ture evolution in the reactor, the heat of reaction, and the specific heat of the reactant.
Another example of this direct technique, which is not based on reactive absorp-
tion, is the stopped-flow method. Barth et al. (1981, 1986) used a stopped-flow
apparatus with light-absorption detection to study CO2-alkanolamine reaction
kinetics. The conversion of the reactant, and hence the reaction kinetics, was
Downloaded by [University of Arizona] at 05:26 27 November 2012

determined by the color change of a pH indicator. Alper (1990) also studied the
kinetics of the reaction between aqueous solutions of CO2 and 2-amino-2-methyl-
1-propanol (AMP) using the stopped-flow method. More recently, Ali et al. (2002)
used this technique to study the kinetics of the reaction of secondary alkanolamines
with CO2 in aqueous solutions. A limitation of such homogeneous experimentation
techniques is that it is not always possible to conduct experiments using high concen-
trations of the solute and reactant.
Although any of the methods described in this section can be used to accurately
measure reaction kinetics, the stirred cell reactor and the laminar jet absorber are the
most versatile and hence most widely used among all the laboratory contactors.

Example of Industrial Importance


The kinetics of CO2 absorption into aqueous DEA solutions has been widely studied
using the different experimental techniques described above (Hikita et al., 1977;
Blanc and Demarais, 1981; Blauwhoff et al., 1984; Savage and Kim, 1985; Barth
et al., 1986; Littel et al., 1992; Rinker et al., 1996). These studies are summarized
in Table VI. DEA is the most popular secondary alkanolamine used for CO2
removal and hence this system is of considerable importance. In this section, the
kinetic behavior of this reaction system is discussed.
The reaction order with respect to CO2 is one (m ¼ 1). When the reaction occurs
entirely in the liquid film and the DEA concentration at the interface is equal to its
concentration in the bulk liquid, Equation (5) can be rewritten as:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
RA ¼ CIA DA kðCB Þn ð22Þ

However, the reaction order with respect to DEA varies, depending on its concen-
tration. Some researchers suggest first- or second-order kinetics while others report
a fractional order between 1 and 2.
Blauwhoff et al. (1984), Mahajani and Joshi (1988), and Rinker et al. (1996)
reviewed the kinetic data of this reaction system and analyzed the information
published. The reaction between CO2 and DEA (denoted here as R2NH, where
R ¼ CH2CH2OH) proceeds through the formation of zwitterion as an intermediate:
k1
CO2 þ R2 NH 30 R2 Nþ HCOO ð23Þ
k1
1560 P. K. Vaidya and E. Y. Kenig

Table VI. Some kinetic studies on the CO2-aqueous DEA reaction system
DEA Order with
concentration respect Experimental
Temp C kmol=m3 to DEA technique Reference

5.8–40.3 0.174–0.719 2 Rapid mixing Hikita et al. (1977)


method
1 Wetted-wall Blanc and Demarais
column (1981)
25 0.39–2.3 Between Stirred cell Blauwhoff et al. (1984)
1 and 2 reactor
50 0.5–2.5 1 Wetted-sphere Savage and Kim (1985)
contactor
Downloaded by [University of Arizona] at 05:26 27 November 2012

25 0.0192–0.0212 1 Stopped-flow Barth et al. (1986)


method
30–60 0.2–3 Between Stirred cell Littel et al. (1992)
1 and 2 reactor
20–70 0.25–2.8 1 Laminar jet Rinker et al. (1996)
absorber

This zwitterion undergoes deprotonation by a base B, thereby resulting in carbamate


formation (Caplow, 1968; Danckwerts, 1979):
kB
R2 Nþ HCOO þ B ! R2 NCOO þ BHþ ð24Þ
Depending on the rate-limiting step (zwitterion formation or deprotonation), the reac-
tion order with respect to DEA is between 1 and 2. Cadours et al. (2007) concluded
that the order is 1 when the gas–liquid contact time is small (wetted-wall column,
wetted-sphere contactor, and laminar jet absorber). In a stirred-cell reactor, where
the contact time is larger, a fractional order between 1 and 2 is found. Hikita et al.
(1977) studied kinetics using the rapid-mixing technique and reported a reaction order
of 2 with respect to DEA. The reaction rate expression and the rate constant differ,
depending on the DEA concentration range studied. Thus, the CO2-aqueous DEA
system exhibits complex kinetic behavior.
From this example, it is clear that different experimental methods may lead
to different estimates of the kinetic parameters for the same reaction system. There-
fore, care has to be taken when interpreting kinetic data and applying it in industrial
practice.

Conclusions
In this work, the methods of determining gas–liquid reaction kinetics are reviewed
and evaluated. The contactors such as stirred cell, wetted-wall column, wetted-
sphere, hemisphere, laminar jet absorber, and disk column that are used for measur-
ing kinetic parameters are described. The design, theory, experimental procedure,
advantages, and disadvantages of each technique are discussed. The homogeneous
experimentation methods such as the rapid mixing method and the stopped-
flow method are also reviewed. Any of these methods can be used for accurately
Gas–Liquid Reaction Kinetics 1561

measuring absorption rates; however, the stirred cell reactor and the laminar jet
absorber belong to the most widely used techniques. The procedure used for the esti-
mation of the kinetic parameters from absorption rate measurements is outlined and
the kinetics of the complex CO2-DEA reaction system representing an industrially
important example is discussed. It is shown that different techniques may lead to
different estimates of the kinetic parameters. Therefore careful interpretation of
the kinetic data is required before its application for industrial cases.

Acknowledgments
PDV is grateful to the Alexander von Humboldt foundation for the financial
support.
Downloaded by [University of Arizona] at 05:26 27 November 2012

Nomenclature
a gas–liquid interfacial area, m2=m3
A solute gas
B reactant in liquid phase
CIA liquid-phase concentration of A at the interface, kmol=m3
CA concentration of A in bulk liquid, kmol=m3
CB concentration of B in bulk liquid, kmol=m3
d diameter of the wetted-wall column, m
dj diameter of the laminar jet, m
DA diffusivity of A in liquid phase, m2=s
DB diffusivity of B in liquid phase, m2=s
E enhancement factor due to chemical reaction
Ei enhancement factor for instantaneous chemical reaction
g gravitational acceleration, m=s2
h height of the wetted-wall column, m
hj height of the laminar jet, m
HA Henry’s coefficient, kmol=(m3 kPa)
Ha Hatta number as defined in Equation (2)
k reaction rate constant as defined by Equation (22)
k1 forward reaction rate constant in Equation (23)
k 1 backward reaction rate constant in Equation (23)
kB reaction rate constant as defined by Equation (24)
kG gas-side mass transfer coefficient, kmol=(m2 s kPa)
kL liquid-side mass transfer coefficient, m=s
kmn reaction rate constant, (m3=kmol)m þ n  1
kov overall reaction rate constant as defined by Equation (15)
L liquid flow rate, m3=s
m reaction order with respect to A
n reaction order with respect to B
nA,G molar holdup of A in gas phase
nA,L molar holdup of A in liquid phase
N speed of agitation, rpm
pA partial pressure of A in bulk gas phase, kPa
pAI partial pressure of A at the gas–liquid interface, kPa
P reactor pressure, kPa
1562 P. K. Vaidya and E. Y. Kenig

r rate of reaction, kmol=(m3 s)


rs radius of wetted sphere, m
R0 gas constant, m3 atm=(kmol K)
RA specific rate of absorption, kmol=(m2 s)
RA average rate of absorption, kmol=s
t time, s
tc contact time, s
T temperature, K
VG volume of gas phase, m3
VL volume of liquid, m3
z stoichiometric reaction coefficient
Greek Letters
Downloaded by [University of Arizona] at 05:26 27 November 2012

a, bi, ci coefficients in Equation (13)


dg gas-film thickness, m
dL liquid-film thickness, m
D liquid-film thickness at the equator of the sphere, m
g viscosity of liquid, kg=(m s)
q density of liquid, kg=m3
n kinematic viscosity of liquid, m2=s
Abbreviations
AEPD 2-amino-2-ethyl-1,3-propanediol
AMP 2-amino-2-methyl-1-propanol
AMPD 2-amino-2-methyl-1,3-propanediol
DEA diethanolamine
DEG diethylene glycol
DGA diglycolamine
DIPA diisopropanolamine
MDEA methyldiethanolamine
MEA monoethanolamine
MOR morpholine
NMP N-methyl-2-pyrrolidone
PE 2-piperidineethanol
PEG polyethyleneglycol
PIP piperazine
TEA triethanolamine

References
Aboudheir, A., Tontiwachwuthikul, P., Chakma, A., and Idem, R. (2003). Kinetics of the
reactive absorption of carbon dioxide in high CO2-loaded, concentrated aqueous mono-
ethanolamine solutions, Chem. Eng. Sci., 58, 5195–5210.
Aboudheir, A., Tontiwachwuthikul, P., Chakma, A., and Idem, R. (2004). Novel design for
the nozzle of a laminar jet absorber, Ind. Eng. Chem. Res., 43, 2568–2574.
Al-Ghawas, H. A., Hagewiesche, D. P., Ruiz-Ibanez, G., and Sandall, O. C. (1989a).
Physicochemical properties important for carbon dioxide absorption in aqueous methyl-
diethanolamine, J. Chem. Eng. Data, 34, 385–391.
Gas–Liquid Reaction Kinetics 1563

Al-Ghawas, H. A., Ruiz-Ibanez, G., and Sandall, O. C. (1989b). Absorption of carbonyl


sulphide in aqueous methyldiethanolamine, Chem. Eng. Sci., 44, 631–639.
Ali, S. H., Merchant, S. Q., and Fahim, M. A. (2002). Reaction kinetics of some secondary
alkanolamines with carbon dioxide in aqueous solutions by stopped flow technique,
Sep. Purif. Technol., 27, 121–136.
Al-Juaied, M. and Rochelle, G. T. (2006). Absorption of CO2 in aqueous blends of diglyco-
lamine and morpholine, Chem. Eng. Sci., 61, 3830–3837.
Alper, E. (1990). Reaction mechanism and kinetics of aqueous solutions of 2-amino-2-
methyl-1-propanol and carbon dioxide, Ind. Eng. Chem. Res., 29, 1725–1728.
Amararene, F. and Bouallou, C. (2004). Kinetics of carbonyl sulfide (COS) absorption with
aqueous solutions of diethanolamine and methyldiethanolamine, Ind. Eng. Chem. Res.,
43, 6136–6141.
Astarita, G., Savage, D. W., and Bisio, A. (1983). Gas Treating with Chemical Solvents, John
Wiley, New York.
Downloaded by [University of Arizona] at 05:26 27 November 2012

Barth, D., Tondre, C., Lappal, G., and Delpuech, J.-J. (1981). Kinetic study of carbon dioxide
reaction with tertiary amines in aqueous solutions, J. Phys. Chem., 85, 3660–3667.
Barth, D., Tondre, C., and Delpuech, J.-J. (1986). Stopped-flow investigations of the reaction
kinetics of carbon dioxide with some primary and secondary alkanolamines in aqueous
solutions, Int. J. Chem. Kinet., 18, 445–457.
Blanc, C. and Demarais, G. (1981). Vitesses de la reaction du CO2 avec la diethanolamine,
Entropie, 102, 53.
Blauwhoff, P. M. M., Versteeg, G. F., and van Swaaij, W. P. M. (1984). A study on the
reaction between CO2 and alkanolamines in aqueous solutions, Chem. Eng. Sci., 39,
207–225.
Cadours, R., Roquet, D., and Perdu, G. (2007). Competitive absorption-desorption of acid
gas into water-DEA solutions, Ind. Eng. Chem. Res., 46, 233–241.
Camacho, F., Sanchez, S., Pacheco, R., Sanchez, A., and D. La Rubia, M. (2005). Absorption
of carbon dioxide at high partial pressures in aqueous solutions of di-isopropanolamine,
Ind. Eng. Chem. Res., 44, 7451–7457.
Caplow, M. (1968). Kinetics of carbamate formation and breakdown, J. Am. Chem. Soc., 90,
6795–6803.
Cullinane, J. T. and Rochelle, G. T. (2006). Kinetics of carbon dioxide absorption into aque-
ous potassium carbonate and piperazine, Ind. Eng. Chem. Res., 46, 2531–2545.
Danckwerts, P. V. (1970). Gas-Liquid Reactions, McGraw-Hill, New York.
Danckwerts, P. V. (1979). The reaction of CO2 with ethanolamines, Chem. Eng. Sci., 34,
443–446.
Davidson, J. F. and Cullen, E. J. (1957). The determination of diffusion coefficients for
sparingly soluble gases in liquids, Trans. Inst. Chem. Eng., 35, 51–60.
Davis, R. A. and Sandall, O. C. (1993). Kinetics of the reaction of carbon dioxide with
secondary amines in polyethylene glycol, Chem. Eng. Sci., 48, 3187–3193.
Dehouche, Z. and Lieto, J. (1995). Modelling and experimental study of key parameters of
absorption on wetted sphere contactor, Chem. Eng. Sci., 50, 2899–2909.
Doraiswamy, L. K. and Sharma, M. M. (1984). Heterogeneous Reactions: Analysis, Examples
and Reactor Design, vol. 2, John Wiley, New York.
Ebrahimi, S., Kleerebezem, R., van Loosedrecht, M. C. M., and Heijnen, J. J. (2003). Kinetics
of the reactive absorption of hydrogen sulfide into aqueous ferric sulphate solutions,
Chem. Eng. Sci., 58, 417–427.
Hartridge, H. and Roughton, F. J. W. (1923). A method of measuring the velocity of very
rapid chemical reactions, Proc. R. Soc. Lond. Ser. A, 104, 376–394.
Higbie, R. (1935). The rate of absorption of a pure gas into a still liquid during short periods
of exposure, Trans. Am. Inst. Chem. Eng., 31, 365–389.
Hikita, H. and Asai, S. (1964). Gas absorption with (m,n)th order irreversible chemical reac-
tion, Int. Chem. Eng., 4, 332–340.
1564 P. K. Vaidya and E. Y. Kenig

Hikita, H., Asai, S., Ishikawa, H., and Honda, M. (1977). The kinetics of reactions of carbon
dioxide with monoethanolamine, diethanolamine and triethanolamine by a rapid mixing
method, Chem. Eng. J., 13, 7–12.
Horng, S.-Y. and Li, M.-H. (2002). Kinetics of absorption of carbon dioxide into aqueous
solutions of monoethanolamine þ triethanolamine, Ind. Eng. Chem. Res., 41, 257–266.
Jamal, A. (2002). Absorption and desorption of CO2 and CO in alkanolamine systems, Ph.D.
diss., University of British Columbia, Vancouver, B. C.
Jamal, A., Meisen, A., and Jim Lim, C. (2006). Kinetics of carbon dioxide absorption and
desorption in aqueous alkanolamine solutions using a novel hemispherical contactor. I:
Experimental apparatus and mathematical modelling, Chem. Eng. Sci., 61, 6571–6589.
Kenig, E. Y., Kucka, L., and Gorak, A. (2003). Rigorous modelling of reactive absorption
processes, Chem. Eng. Technol., 26, 631–646.
Kucka, L., Kenig, E. Y., and Gorak, A. (2002). Kinetics of the gas-liquid reaction between
carbon dioxide and hydroxide ions, Ind. Eng. Chem. Res., 41, 5952–5957.
Downloaded by [University of Arizona] at 05:26 27 November 2012

Kucka, L., Richter, J., Kenig, E. Y., and Gorak, A. (2003). Determination of gas-liquid reac-
tion kinetics with a stirred cell reactor, Sep. Purif. Technol., 31, 163–175.
Lee, S., Song, H.-J., Maken, S., and Park, J.-W. (2007). Kinetics of CO2 absorption in aque-
ous sodium glycinate solutions, Ind. Eng. Chem. Res., 46, 1578–1583.
Lewis, W. K. and Whitman W. G. (1924). Principles of gas absorption, Ind. Eng. Chem., 16,
1215–1220.
Liao, C.-H. and Li, M.-H. (2002). Kinetics of absorption of carbon dioxide into aqueous solu-
tions of monoethanolamine þ N-methyldiethanolamine, Chem. Eng. Sci., 57, 4569–4582.
Littel, R. J., Versteeg, G. F., and van Swaaij, W. P. M. (1992). Kinetics of CO2 with primary
and secondary amines in aqueous solutions. II: Influence of temperature on zwitterion
formation and deprotonation rates, Chem. Eng. Sci., 47, 2037–2045.
Long, X.-L., Xiao, W.-D., and Yuan, W.-K. (2005). Kinetics of gas-liquid reaction between
NO and Co(en)33 þ , Ind. Eng. Chem. Res., 44, 4200–4205.
Lynn, S., Straatemeier, J. R., and Kramers, H. (1955). Absorption studies in the light of the
penetration theory. III: Absorption by wetted spheres, singly and in columns, Chem. Eng.
Sci., 4, 63–67.
Mahajani, V. V. and Joshi, J. B. (1988). Kinetics of reactions between CO2 and alkanolamines,
Gas Sep. Purif., 2, 50–64.
Ramachandran, N., Aboudheir, A., Idem, R., and Tontiwachwuthikul, P. (2006). Kinetics of
the absorption of CO2 into mixed aqueous loaded solutions of monoethanolamine and
methyldiethanolamine, Ind. Eng. Chem. Res., 45, 2608–2616.
Rinker, E. B., Ashour, S. S., and Sandall, O. C. (1995). Kinetics and modelling of carbon diox-
ide absorption into aqueous solutions of N-methyldiethanolamine, Chem. Eng. Sci., 50,
755–768.
Rinker, E. B., Ashour, S. S., and Sandall, O. C. (1996). Kinetics and modeling of carbon
dioxide absorption into aqueous solutions of diethanolamine, Ind. Eng. Chem. Res., 35,
1107–1114.
Rinker, E. B., Ashour, S. S., and Sandall, O. C. (2000). Absorption of carbon dioxide into
aqueous blends of diethanolamine and methyldiethanolamine, Ind. Eng. Chem. Res.,
39, 4346–4356.
Savage, D. W. and Kim, C. J. (1985). Chemical kinetics of carbon dioxide reactions with
diethanolamine and diisopropanolamine in aqueous solutions, AIChE J., 31, 296–301.
Savage, D. W., Astarita, G., and Joshi, S. (1980). Chemical absorption and desorption of
carbon dioxide from hot carbonate solutions, Chem. Eng. Sci., 35, 1513–1522.
Seo, D. J. and Hong, W. H. (2000). Effect of piperazine on the kinetics of carbon dioxide with
aqueous solutions of 2-amino-2-methyl-1-propanol, Ind. Eng. Chem. Res., 39, 2062–2067.
Shen, K.-P., Li, M.-H., and Yih, S. M. (1991). Kinetics of carbon dioxide reaction with
sterically hindered 2-piperidineethanol aqueous solutions, Ind. Eng. Chem. Res., 30,
1811–1813.
Gas–Liquid Reaction Kinetics 1565

Stephens, E. J. and Morris, G. A. (1951). Determination of liquid-film absorption coefficients,


Chem. Eng. Prog., 5, 232–242.
Sun, W.-C., Yong, C.-B., and Li, M.-H. (2005). Kinetics of the absorption of carbon dioxide
into mixed aqueous solutions of 2-amino-2-methyl-1-propanol and piperazine, Chem.
Eng. Sci., 60, 503–516.
Tomcej, R. A. and Otto, F. D. (1989). Absorption of CO2 and N2O into aqueous solutions of
methyldiethanolamine, AIChE J., 35, 861–864.
Tseng, P. C., Ho, W. S., and Savage, D. W. (1988). Carbon dioxide absorption into promoted
carbonate solutions, AIChE J., 34, 922–931.
Vaidya, P. D. and Mahajani, V. V. (2005). Kinetics of the reaction of CO2 with aqueous for-
mulated solution containing monoethanolamine, N-methyl-2-pyrrolidone and diethylene
glycol, Ind. Eng. Chem. Res., 44, 1868–1873.
van Swaaij, W. P. M. and Versteeg, G. F. (1992). Mass transfer accompanied with complex
reversible reactions in gas-liquid systems: An overview, Chem. Eng. Sci., 47, 3181–3195.
Downloaded by [University of Arizona] at 05:26 27 November 2012

Xiao, J., Li, C.-W., and Li, M.-H. (2000). Kinetics of absorption of carbon dioxide into aque-
ous solutions of 2-amino-2-methyl-1-propanol þ monoethanolamine, Chem. Eng. Sci., 55,
161–175.
Xu, G.-W., Zhang, C.-F., Qin, S.-J., and Wang, Y.-W. (1992). Kinetics study on absorption of
carbon dioxide into solutions of activated methyldiethanolamine, Ind. Eng. Chem. Res.,
31, 921–927.
Yih, S.-M. and Shen, K.-P. (1988). Kinetics of carbon dioxide reaction with sterically hindered
2-amino-2-methyl-1-propanol aqueous solutions, Ind. Eng. Chem. Res., 27, 2237–2241.
Yoon, S. J., Lee, H., Yoon, J.-H., Shim, J.-G., Lee, J. K., Min, B.-Y., and Eum, H.-M. (2002).
Kinetics of absorption of carbon dioxide into aqueous 2-amino-2-ethyl-1,3-propanediol
solutions, Ind. Eng. Chem. Res., 41, 3651–3656.
Yoon, J.-H., Baek, J.-I., Yamamoto, Y., Komai, T., and Kawamura, T. (2003). Kinetics of
removal of carbon dioxide by aqueous 2-amino-2-methyl-1,3-propanediol, Chem. Eng.
Sci., 58, 5229–5237.
Zhang, X., Zhang, C.-F., Qin, S.-J., and Zheng, Z.-S. (2001). A kinetics study on the absorp-
tion of carbon dioxide into a mixed aqueous solution of methyldiethanolamine and
piperazine, Ind. Eng. Chem. Res., 40, 3785–3791.
Zhang, X., Zhang, C.-F., and Liu, Y. (2002). Kinetics of absorption of CO2 into aqueous
solution of MDEA blended with DEA, Ind. Eng. Chem. Res., 41, 1135–1141.

You might also like