You are on page 1of 8

Article

pubs.acs.org/IECR

Reactive Absorption of Carbon Dioxide in LProlinate Salt Solutions


Magdalena E. Majchrowicz,* Sascha Kersten, and Wim Brilman
SPT Group, Faculty of Science and Technology, University of Twente, PO Box 217, 7500 AE Enschede, The Netherlands
S Supporting Information
*

ABSTRACT: Aqueous amino acid salt solutions are seen as more sustainable alternative CO2 solvents compared to
conventional alkanolamines. The absorption of CO2 into aqueous solutions of potassium L-prolinate, over the temperature range
of 290303 K, was studied using a stirred cell. To compare the eect of potassium versus sodium as counterion, the absorption
rates of CO2 in sodium L-prolinate solutions were also determined. The amino acid salt concentration was varied between 0.5 and
3 mol dm3. Physicochemical properties such as density, viscosity, and physical solubility of N2O, required in the interpretation
of absorption rate experiments, were determined separately. The obtained experimental reactive absorption uxes were
interpreted, using the pseudo-rst-order approach, into intrinsic reaction kinetics. The potassium-based solvent showed, on
average, a 32% higher reactivity toward CO2 than the sodium equivalent. The kinetic data were correlated in a power-law reaction
rate expression. The reaction order with respect to the amino acid was found to be between 1.40 and 1.44 for both L-prolinate
salts, and the second order kinetic rate constant, k2, was calculated for the potassium salt to be 93.7 103 dm3mol1s1 at 298 K
with an activation energy of 43.3 kJmol1. L-Prolinate salts show higher chemical reactivity toward CO2 than most of the aminoalcohols and amino acid salts.
Potassium salt of glycine was studied also by Portugal et al.8
who measured the absorption rates of CO2 at temperatures
from 293 to 313 K, over the solution concentration range of
0.13 moldm3. Portugal et al.9 determined the reaction
kinetics of CO2 in potassium threonate aqueous solution at
concentration ranging from 0.13 moldm3 and temperatures
from 293 to 313 K. Van Holst et al.7 related the overall kinetic
constant with pKa for several amino acid salts by means of a
Brnsted plot. Potassium salts of 6-aminohexanoic acid, alanine, L-arginine, L-glutamic acid, DL-methionine, L-proline
and sarcosine were considered. It was argued that a high
reaction rate is important for reducing the size and hence
capital costs of the absorber, while relatively low pKa is of
importance for minimizing the energy input in the desorber.
On the basis of this criterion, alkaline salts of L-proline and
sarcosine were identied as promising amino acid salt solvents
for CO2 capture. Simons et al.10 studied the absorption rate of
CO2 in aqueous potassium sarcosinate solutions. A change of
the reaction rate constant with temperature (T = 298308 K,
Cm = 1.5 M), amino acid salt concentration (T = 298 K, Cm =
0.53.8 M) and CO2 liquid loading ( = 00.5 mol CO2 per
mol solution, T = 298 K, Cm = 1.45 M) was reported. Paul and
Thomsen11 studied the absorption rate of CO2 in aqueous
potassium L-prolinate solutions at 303, 313, and 323 K within
the salt concentration range of 0.53 moldm3.
In the present work, the absorption of CO2 into aqueous
solutions of potassium L-prolinate over the temperature range
of 290303 K using a stirred cell apparatus was studied. The
absorption rates of CO2 were also measured for the sodium salt
at 298 K. The amino acid salt concentration varied between 0.5

1. INTRODUCTION
Absorption by a chemical reaction is a common process in the
chemical industry and it is used, among others, in the treatment
of the industrial gas streams containing acid gases like H2S,
NOx, and CO2. In these gas treating processes, aqueous
solutions of amino-alcohols (MEA, DEA, MDEA, AMP, their
mixtures with or without activators, etc.) are commonly used.
These solvents, however, exhibit several drawbacks such as the
limited cyclic CO2 working capacity, necessity of the addition of
corrosion inhibitors to prevent high equipment corrosion rates,
high energy consumption for regeneration, solvent losses by
evaporation, and thermal- and oxidative degradation.1 Aqueous
solutions of amino acid salts can be an interesting alternative
for the commercially used amine-based absorbents. In general,
amino acid salts are characterized by more favorable properties
like low volatility (due to their ionic nature) and higher surface
tension (important for membrane gas absorption ( MGA))2,3
and a high level of biodegradability in combination with low
toxicity.4 At the same time they maintain (due to similar
functionality) a comparable or even higher chemical reactivity
toward CO2 as compared to amino-alcohols.57 Owing to these
favorable properties, amino acid salts have been deployed in the
past as rate promoters in carbonate solutions in the
GiammarcoVetrocoke process and for the selective acid gas
removal from industrial gas streams as for example in the
Alkacid process by BASF.1 In the recent years, Siemens AG has
developed a proprietary postcombustion carbon capture
technology, POSTCAP, based on aqueous amino-acid salt
solutions.
Knowledge of the reaction kinetics between CO2 and amino
acid salts is essential for accurate design and simulation of the
CO2 absorption process. Kumar et al.6 measured the reaction
kinetics of CO2 in potassium salts of taurine at temperatures of
285, 295, and 305 K and potassium glycinate at 295 K, and over
the solution concentration ranging from 0.1 to 4 moldm3.
2014 American Chemical Society

Received:
Revised:
Accepted:
Published:
11460

March 15, 2014


June 22, 2014
June 22, 2014
June 22, 2014
dx.doi.org/10.1021/ie501083v | Ind. Eng. Chem. Res. 2014, 53, 1146011467

Industrial & Engineering Chemistry Research

Article

and 3 moldm3. After validating that the absorption experiments were conducted in the so-called pseudo-rst-order
regime, the reaction rate constants between L-prolinate salts
and CO2 were extracted from the obtained experimental data
using this approach. Complementary to the kinetic data, the
physicochemical properties such as density, viscosity and
physical solubility of N2O in L-prolinate salts were determined.
These data are reported in the Supporting Information.
Liquid bulk concentrations of all chemical species are
required for kinetic analysis. A vaporliquid equilibrium
(VLE) model to estimate the CO2 partial pressure and the
liquid bulk concentrations of all chemical species present in the
solution was earlier described in Majchrowicz and Brilman.12
Using the experimental data from the present work, this model
could be further extended to a comprehensive absorption rate/
kinetics model for interpreting the absorption data of CO2 in
aqueous amino acid salt solutions.

amines.16 Both asymptotic options seem plausible for alkaline


salts of L-proline. Similarly to DEA, this amino acid has a
secondary amino functional group. On the other hand, Lprolinate has higher pKa than MEA which should result in a
high deprotonation rate constant kB.

kB

k H2O

k 2[RNH][CO2 ]
1+

k 1
kB[B]

[RNH][CO2 ]
1
k2

k 1
1
k 2 kB[B]

(8)

For amino acid salts, the contribution of reactions 7 and 8 to


the overall rate of CO2 absorption is usually neglected.69 If the
reaction between CO2 and OH ions is taken into account then
the overall reaction kinetic constant can be expressed by eq 9,
where kapp is the apparent rate constant for reaction 7. The
concentration of OH is obtained by pH measurements. The
rate constant kOH is correlated to temperature and ionic
strength by eqs 10 and 11.17

(1)

kov = kapp + k OH[OH]

(9)

k OH
= aI bI 2

k OH

(10)

log

2382
(11)
T

Here, kOH is the rate constant of reaction 7 in an innitely


diluted solution, I is the solution ionic strength and constants a
and b for dierent counterions are given in Table 1.

logkOH
= 11.895

(3)

Table 1. Constants in Equation 10 (Pohorecki and


Moniuk17)

Assuming a quasi-steady state condition for the zwitterion


concentration and a rst order behavior of CO2, the overall
forward rate of reaction of CO2 with amino acid salt can be
expressed with eq 4.14
rCO2 =

(7)

CO2 + H 2O HCO3 + H+

(2)

RNH+COO + B RNCOO + BH+

(6)

k OH

k2

k 1

kB[B]
rCO2 = k 2[CO2 ][RNH]

k 1

CO2 + OH HCO3

The zwitterion mechanism, originally proposed by Caplow,13


is generally applied to model the CO2 absorption in aqueous
solutions of amino-alcohols or amino acid salts (AAS). This
reaction mechanism involves the formation of the intermediate,
a zwitterion (eq 2), followed by the removal of a proton by a
base B (eq 3). The deprotonation reaction can be considered to
be irreversible (equilibrium predominantly to the right-hand
side of the reaction).
CO2 + RNH XooY RNH+COO

(5)

Besides the carbamate formation from amino acid salt by


reaction 2, other reactions can occur in the CO2AASH2O
systems.

2. REACTION MECHANISM
In aqueous solutions, CO2 reacts with an alkaline salt of
(primary or secondary) amino acid to the corresponding
carbamate according to the overall reaction 1 (here RNH
corresponds to L-prolinate with nonreacting part RC4H7
COO).
2RNH + CO2 RNH+2 + RNCOO

rCO2 = k 2[CO2 ][RNH]

system

CO2KOH
CO2NaOH

0.287
0.211

0.013
0.016

= kapp[CO2 ]

3. MASS TRANSFER
Generally, the absorption rate of CO2 in a reactive liquid, in the
absence of gas-phase limitations and assuming ideal gas
behavior under the low pressures applied in the experiments,
is given by eq 12.18,19
mCO2pCO
2
JCO = kLECO2
2
(12)
RT
where JCO2 is the molar ow of CO2 entering the liquid, kL is
the liquid side mass transfer coecient, pCO2 is partial pressure
of CO2 in the reactor, mCO2 is the distribution coecient of
CO2 between the gas and liquid phase and ECO2 is the
enhancement factor. The parameter mCO2 describes the physical
solubility of CO2 in the solution and is equal to (RT) divided
by the Henry coecient He. It is dened by mCO2 = Cl,i/Cg,i, in

(4)

where kb[B] designates the contribution of all bases present


in the solution for the proton removal step. In lean aqueous
solutions, the liquid phase reactant, H2O and OH species can
act as bases.15
Limiting conditions will lead to more simplied reaction rate
expressions.6 If the deprotonation of the zwitterion is relatively
fast compared to the reversion rate of CO2 and amine, hence
k1/ kb[B] 1, the kinetic rate expression reduces to simple
second order kinetics as found for primary alkanolamines such
as MEA (see eq 5). If, on the other hand, k1/ kb[B] 1
then the kinetic rate expression reduces to eq 6. In this case the
reaction order depends on the contribution of the individual
bases to the deprotonation of the zwitterion. This expression
accounts as well for the reaction order shift with changing
amine concentration as found for many secondary alkanol11461

dx.doi.org/10.1021/ie501083v | Ind. Eng. Chem. Res. 2014, 53, 1146011467

Industrial & Engineering Chemistry Research

Article

proposed by Weisenberger and Schumpe27 for electrolyte


solutions was applied to correlate the salting out eect.
mH O
log 2 = K Cm
m
(17)

which Cg,i is equal to the molar concentration of CO2 in the gas


phase and Cl,I is the corresponding physical solubility of CO2 in
the liquid phase. The enhancement factor is the ratio of the
chemical absorption ux to the absorption ux in the absence
of a chemical reaction. The enhancement factor is a function of
the Hatta number Ha and is limited by the innite
enhancement factor E. The Hatta number is dened by eq
13 and relates the maximum conversion rate within the mass
transfer zone to the diusional transport through the mass
transfer zone.20 The innite enhancement factor is determined
by the diusion of reactants toward the reaction zone and thus
corresponds to an instantaneous reaction. Its value can be
estimated, according to the penetration theory, by eq 14.21 It
should be noted, however, that this equation is valid only for
irreversible reactions. In the present work, owing to a high
value for the equilibrium constant of reaction 1 (Majchrowicz
and Brilman12) and negligible liquid CO2 loadings present in
the experiments, this reaction is regarded as essentially
irreversible.
Ha =

E =

Here mH2O and m denote the gas solubility in pure water and
the salt solution, respectively, Cm is the molar concentration of
the salts. For a single salt, the Sechenov constant K is then
given by the following relation:
K=

hg = hg,0 + h T(T 298.15 K)

(14)

In the above equations, kov is the overall reaction kinetic


constant, DCO2 is the diusion coecient of CO2 in a solution,
DAm is the diusion coecient of the liquid phase reactant in a
solution and vAAS is the stoichiometric coecient that in
analogy to alkanolamines has a value of 2.10,22,23
Three absorption regimes can be distinguished depending on
the absolute value of the Hatta number and the Ha/E
ratio.8,9,18,19 Absorption taking place in the fast pseudo-rstorder (PFO) reaction regime can be assumed if the following
condition is satised:
3 < E = Ha E

DN2O
DN2O,H2O

kovDCO2

H O
= 2 = constant

(20)

where usually 0.5 < < 1. The N2OCO2 analogy was further
used to estimate the diusivity of CO2 in L-prolinate
solutions.29 The value of 0.6 was used to estimate the
diusion coecient of N2O in aqueous solutions of L-prolinate
(Paul and Thomsen11). This value was used also for MEA
(Portugal et al.8). For potassium salts of L-proline, the
viscosities reported by van Holst et al.30 were used. For
sodium salts, the viscosities obtained in the present work and
reported in the Supporting Information were applied. The
viscosities of water H2O were taken from Popiel and
Wojtkowiak.31 The diusivity of CO2 and N2O in water were
obtained from Versteeg and van Swaaij.28 The diusion
coecients of L-prolinate in aqueous solution, as required in
eq 14, were taken from the work by Hamborg et al.32

(15)

In this absorption regime, the enhancement factor equals the


Ha-number, which makes it possible to obtain the kinetics for
the forward reaction, starting with unloaded solutions.
Consequently, eq 12 changes to
JCO =

(19)

Here hi and hg are the ion and gas specic parameters,


respectively, and ni is the index of ion i.
The gas-specic parameters hg and the ion-specic constants
hi (of ions other than L-prolinate) were obtained from
Weisenberger and Schumpe.27 The L-prolinate ion-specic
parameter was obtained (as a function of temperature) from
the measured N2O solubility in this solvent, see the Supporting
Information. The distribution coecient of CO2 and N2O in
water were obtained from Versteeg and van Swaaij.28 The
diusivity of N2O in the absorption liquids was calculated using
the StokesEinstein relation:

(13)

DCO2,
D
[AAS]RT
1 + AAS
DAAS
DCO2 AmPCO2mCO2

(18)

with

kovDCO2
kL

(hi + hg)ni

mCO2pCO

RT

(16)

4. EXPERIMENTAL SECTION
The CO2 [124-38-9] gas used was obtained from Linde Gas. Lproline ((S)-Pyrrolidine-2-carboxylic acid, C5H9NO2, 98.5%)
[147-85-3], potassium hydroxide (85.0%; with the actual
purity determined accurately by titration) [1310-58-3], sodium
hydroxide (98.0%) [1310-65-2], MEA (2-aminoethanol,
C2H7NO, 99.0%) [141-43-5] were purchased from SigmaAldrich and were used without further purication. Aqueous Lprolinate solutions were prepared by neutralizing the amino
acid dissolved in deionized, double-distilled water with an
equimolar quantity of an alkaline hydroxide. The amino acid
salt concentration was estimated potentiometrically by titrating
with a standard 1.000 moldm3 HCl solution [7647-01-0].
A schematic representation of the gasliquid contactor used
to study the reaction kinetics is given in Figure 1 and is
identical to the one used by Simons et al.10 The experimental
setup consisted essentially of the thermostated cell operated

The instantaneous reaction regime refers to conditions for


which E Ha. In this regime, the reaction is said to be
instantaneous with respect to mass transfer, and determination
of the reaction kinetics from the absorption rate experiments is
not possible. In between the above two reaction regimes, there
is an intermediate regime for which the enhancement factor
can be approximated as a function of Ha and E.24,25
Physicochemical properties such as physical solubility of CO2
and diusion coecients are needed to obtain the kinetic rate
coecient estimates. Since CO2 reacts with amino acid salts,
the physical solubility of CO2 in these solvents is not
measurable and needs to be estimated indirectly from solubility
experiments with N2O. This nonreactive gas is widely used for
this purpose because it has a molecular conguration, volume,
and electronic structure similar to that of CO2.26 Additionally,
for gas solubility in aqueous L-prolinate solutions, the model
11462

dx.doi.org/10.1021/ie501083v | Ind. Eng. Chem. Res. 2014, 53, 1146011467

Industrial & Engineering Chemistry Research

Article

solutions of monoethanolamine (MEA) at 288, 298, and 303 K


was measured. Supporting Information, Table S6 compares the
observed k2 values with those obtained using the rate constant
expression by Versteeg et al.,16 k2 = 4.4 108 exp(5400/T),
and these are considered to be in good agreement.
The absorption rate of CO2 in aqueous potassium L-prolinate
was investigated at temperatures of 290, 298, and 303 K, over
the solution concentration range of 0.5 to 3 moldm3. The
reaction kinetics of CO2 in sodium L-prolinate was also
determined from absorption experiments at 298 K for the same
solution concentration range. In the determination of the
kinetic rates, only the data at suciently low CO2 partial
pressure, where a linear relation exists between the
experimentally observed ux and the CO2 partial pressure
(hence, where E = Ha) were used; as this constant slope
indicates pseudo-rst-order behavior. The results of the kinetic
experiments are presented in Supporting Information, Tables
S7 and S8 for potassium and sodium salts, respectively. For
both salts of L-proline the condition described by eq 15 is
satised and the ratio between the innite enhancement factor
and the Ha number is about a factor of 4 and higher. In all
kinetic measurements, the values of the CO2 loading of the
solution (in mol CO2 per mol amine-group in a solution) at
the end of the experiments were always lower than 5 103.
This should ensure no precipitation in the considered
absorption systems and no signicant decrease of the free
amine concentration during the experiment due to CO2
absorption.12,33 Also, the contribution of kOH[OH] to kov is
relatively small, with the maximum deviation of 5.7% for all
experiments carried out and in most cases less than 1%. This
limited eect of OH to kov has been reported also for the most
common weakly basic amines like MEA, DEA, glycine, taurine,
sarcosine.
Figure 2 compares the overall rate constants of potassium Lprolinate with other amino acid salts, MEA, DEA, and
piperazine at 298 K. For taurine, the kinetic data are reported
at 295 K. Generally, the (secondary) ring-structure amines such
as L-proline and piperazine absorb CO2 at considerable rates.
Among amino acid salts, the reaction kinetics is the lowest for
threonate and taurate, and is similar to that of MEA, whereas Lprolinate shows comparable (high) reactivity toward CO2 to
sarcosinate and glycinate. Lower absorption rates found for
threonate can be explained by the steric hindrance caused by
the hydroxyethylene group connected to the carbon.9
Figure 3 shows the apparent rate constants of potassium Lprolinate for 290, 298, and 303 K. For the three temperatures
studied, a clear temperature dependence of the kapp can be
observed. In Figure 4 the apparent rate constants of potassium
salt of L-proline are plotted together with the kapp data obtained
for the sodium salt at 298 K. A loglog plot of the apparent
rate constant versus the CO2 absorbent concentration is
generally used to identify the mechanism of the reaction
between amine and CO2. The reaction order n with respect to
amine is given by the slope of the graph. For L-prolinate, this
yields the n value of 1.401.44 for potassium- and sodium salts.
Paul and Thomsen11 investigated the reaction kinetics of
potassium L-prolinate over the temperature range of 303323
K using a wetted wall column absorber and reported the
reaction order in amino acid to be in the range of 1.361.40.
Their kinetic data reported for 303 K are included in Figure 3.
In a screening study, van Holst et al.7 studied the kinetics of
CO2-aqueous potassium L-prolinate at 298 K in a stirred cell
reactor using apparatus similar to that used in the present work.

Figure 1. Schematic diagram of the experimental setup used to


determine the reaction kinetics.

with a smooth, horizontal gasliquid interface and a calibrated


gas supply vessel. Two separate stirrers were used in the reactor
to agitate the gas and liquid phases. The reactor was operated
batch-wise with respect to the liquid phase and semicontinuous with respect to the gas phase. To allow for this
mode of operation, the reactor was connected to the gas vessel
via a constant-pressure regulator. In a typical experiment, a
known amount of a solvent solution (0.60 dm3) was charged
to the reactor and shortly degassed under vacuum. After that,
the vaporliquid equilibrium was allowed to establish in the
reactor at the desired temperature. This was repeated until the
pressure did not change anymore and the vapor pressure of the
solvent was reached. This pressure was recorded as the vapor
pressure, Pvap. The pressure controller was set to the desired
(total) pressure, Ptot, and pure CO2 from the gas vessel was
introduced into the reactor. Next, the stirrers in both phases
were turned on and the pressure in the gas supply vessel was
monitored for around 300 s. The dynamic pressure in the gas
vessel (dPgv/dt) was recorded with a pressure transducer
connected to a computer. The pressure in the reactor was also
continuously recorded, with an accuracy of 0.07 mbar, to verify
whether the CO2 partial pressure (PCO2 = Ptot Pvap) remained
constant. The pressure decrease in the gas vessel was due to the
CO2 absorption in the liquid and can be related to the kinetics
by eq 22; if the experiments are carried out in the pseudo-rst
order absorption regime.
JCO A =
2

dPgv Vgv
dt RTgv

(22)

Here A is the geometric gasliquid interfacial area, Vgv is the


volume and Tgv is the temperature of the gas vessel. In this
reaction regime, the CO2 absorption rate is independent of the
liquid side mass transfer coecient and the generic CO2
absorption rate (JCO2A) is given as (Kumar et al.;6 Simons et
al.10):
PCO2
JCO A = ( kovDCO2 )mCO2
A
2
RT

(23)

5. RESULTS AND DISCUSSION


The physicochemical properties required in the interpretation
of absorption rate experiments for alkaline salts of L-proline can
be found in the Supporting Information.
For validation of the experimental setup and procedures
followed, the absorption rate of CO2 in aqueous 2 moldm3
11463

dx.doi.org/10.1021/ie501083v | Ind. Eng. Chem. Res. 2014, 53, 1146011467

Industrial & Engineering Chemistry Research

Article

Figure 4. Apparent kinetic constant of alkaline salts of L-proline at 298


K: , potassium salt; , sodium salt.

Figure 2. Overall kinetic constant for potassium L-prolinate compared


to other CO2 absorbents at 298 K: , potassium L-prolinate, this work;
, potassium sarcosinate, Simons et al.;10 , potassium glycinate,
Portugal et al.;8 , potassium threonate, Portugal et al.;9 , potassium
taurate, Kumar et al.;6 +, piperazine, Derks et al.;19 , MEA, Laddha
and Danckwerts;34 , DEA, Versteeg and Oyevaar.35

well with our results, the estimates of van Holst et al.7 diverge
from our data by +20.5% for 0.5 moldm3 solution and by
30.5% for the highest solution concentration. It should be
noted, however, that van Holst et al.7 already reported a
standard deviation of 36.0% between their experimental and
calculated values of kapp. The obtained fractional reaction orders
(n = 1.401.44) have also been observed for reactions between
CO2 and secondary amines.14,16,35 Alkaline salts of L-proline
also contain a secondary amino group which in this respect
exhibits the same behavior.
Pohorecki and Moniuk17 showed that for the reaction
between CO2 and hydroxyl ions in electrolyte solutions the
kinetic rate constant depends not only on the reactants
concentration but also on the type of ions present in a solution.
Assumed rst order in the hydroxyl ion and CO2, the kinetic
rate constant was found to vary signicantly with ionic strength.
From their data, the reaction in the potassium hydroxide
solutions was around 30% faster at 1.5 M and 80% faster in 3 M
solutions of potassium hydroxide when compared to that of
sodium hydroxide at the same molar concentration. In this
study, for L-prolinate salts, the overall reaction rate also shows a
dependence on the type of counterion present in a solution
(see Figure 4). The reaction order n with respect to the
prolinate ion concentration is the same for both types of
solution, n is around 1.4. With this, the overall kinetic rate
constant kov for potassium L-prolinate is, on average, 32%
higher when compared to the sodium equivalent and, from the
limited set of data, now independent of the prolinate
concentration. Similarly to hydroxide solutions, the use of the
activity-based kinetic rate expressions, rather than the traditional concentration-based approach, might further explain the
dierence in the rate constants observed.36 To realize this,
however, the individual activity coecients for each of the
reaction components of interest would need to be determined,
which is beyond the current scope.

Figure 3. Apparent kinetic constant of potassium L-prolinate as a


function of temperature and solution concentration. , 290 K; , 298
K; , 303 K; this work; +; 298 K, van Holst et al.;7 , 303 K, Paul and
Thomsen.11

They gave the reaction rate expression with the reaction order
in amino acid of 1.10. The kapp values for temperature of 298 K
obtained by this correlation are also plotted in Figure 3.
Whereas the data reported by Paul and Thomsen11 agree very
11464

dx.doi.org/10.1021/ie501083v | Ind. Eng. Chem. Res. 2014, 53, 1146011467

Industrial & Engineering Chemistry Research

Article

There are two limiting cases in the zwitterion mechanism.


When the zwitterion formation reaction is rate limiting, the
reaction rate appears to be rst-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.14,28 On the other
hand, when the zwitterion deprotonation reactions are rate
determining, the overall reaction rate appears to have a
fractional order between 1 and 2 in the amine concentration, as
was observed in the present study. The kapp data from the
present work for potassium and sodium salts of L-proline were
regressed to the reaction rate expression, eq 4, by means of a
LevenbergMarquardt tting procedure. As earlier indicated,
the contribution of OH ions to deprotonation of the
zwitterion was not signicant and was therefore left out from
eq 4. The tted zwitterion rate coecients are listed in
Supporting Information, Table S9. For potassium L-prolinate
and at the system temperature of 303 K, the second order rate
constant k2 from the present work is in good agreement (4.8%
higher), compared to the value given by Paul and Thomsen11
(k2 = 11.89 104 dm3mol1s1).
The rate coecient estimates for the potassium salt of Lproline were tted as a function of temperature by the
following Arrhenius equations:
43330

k 2 = 3.69 10 exp
RT

(24)

kAAS
24400

= 8.78 106exp
RT
k 1

(25)

14320

= 8.65 106exp
RT

(26)

Figure 5. Arrhenius plot of k2 for aqueous potassium L-prolinate


solution. , calculated with eq 24.

12

k H 2O
k 1

The average deviations for k2, kAAS/k1, kH2O/k1 calculated


from correlations 2426 are 0.1%, 1.7%, and 1.2%, respectively.
The calculated rates of absorption using eqs 2426 are
included in Supporting Information, Table S7. A good
agreement between the predicted and experimental JCO2 values,
with the average absolute deviation of 2.86%, can be found. The
activation energy for k2 is calculated to be 43.3 kJ mol1; which
is reasonably in line with the value of 36.5 kJ mol1 given by
Paul and Thomsen11 for the temperature range of 303323 K.
Arrhenius plots for the zwitterion mechanism rate constants are
shown in Figures 5 and 6. Of the zwitterion rate constants, only
k2 is found to be strongly temperature dependent, while the
ratios kAAS/k1 and kH2O/k1 are found to be less sensitive to
temperature. For temperature of 298 K and 2 moldm3
solution, the contribution of water to the zwitterion
deprotonation rate is found to be 72.5% and 72.7%,
respectively, for potassium and sodium salts. This indicates
that water contributes signicantly to the deprotonation even at
moderately high amino acid salt concentration. Considering the
results obtained in Supporting Information, Table S9, with
deprotonation of the zwitterion being rate determining,
reinterpretation of the kinetic data according to the molecular
mechanism (Crooks and Donnellan37) could be considered.

Figure 6. Arrhenius plot of kinetic constants ratios for aqueous


potassium L-prolinate solution: , kH2O/k1; , kAAS/k1; ,
calculated with eqs 25 and 26.

sodium L-prolinate at 298 K. The amino acid salt concentration


was in the range of 0.5 and 3 moldm3. The kinetic data were
found to be consistent with the zwitterion mechanism with
deprotonation being the rate-determining step. The reaction
order with respect to the amino acid, assuming power-law
kinetics, was found to be between 1.40 and 1.44 for both
considered salts, indicating signicant contributions from both
L-prolinate and water to the zwitterion deprotonation. The

6. CONCLUSIONS
The kinetics of the reaction between CO2 and potassium Lprolinate was estimated from absorption data obtained in a
stirred cell over the temperature range of 290303 K and for
11465

dx.doi.org/10.1021/ie501083v | Ind. Eng. Chem. Res. 2014, 53, 1146011467

Industrial & Engineering Chemistry Research

Article

potassium-based solvent showed, on average, a 32% higher


reactivity toward CO2 than the sodium-based solutions and has
a higher reactivity toward CO2 than most of amino-alcohols
and amino acid salts. Potassium prolinate may be therefore
attractive in the bulk removal of CO2, either as solvent or as
activator to other high capacity, low energy CO2 capture
solutions.

REFERENCES

(1) Kohl, A. L.; Nielsen. R. B. Gas Purication, 5th ed.; Gulf


Publishing Company: Houston, TX, 1997.
(2) Kumar, P. S.; Hogendoorn, J. A.; Feron, P. H.; Versteeg, G. F.
Density, Viscosity, Solubility, and Diffusivity of N2O in Aqueous
Amino Acid Salt Solutions. J. Chem. Eng. Data 2001, 46, 1357.
(3) Simons, K.; Nijmeijer, K.; Mengers, H.; Brilman, W.; Wessling,
M. Highly Selective Amino Acid Salt Solutions as Absorption Liquid
for CO2 Capture in GasLiquid Membrane Contactors. ChemSusChem 2010, 3, 939.
(4) Shao, R.; Stangeland, A. Amines Used in CO2 CaptureHealth
and Environmental Impacts; The Bellona Foundation: Oslo, Norway,
2009
(5) Hook, R. J. An Investigation of Some Sterically Hindered Amines
as Potential Carbon Dioxide Scrubbing Compounds. Ind. Eng. Chem.
Res. 1997, 36, 1779.
(6) Kumar, P. S.; Hogendoorn, J. A.; Feron, P. H.; Versteeg, G. F.
Kinetics of the Reaction of CO2 with Aqueous Potassium Salt of
Taurine and Glycine. AIChE J. 2003, 49, 203.
(7) Van Holst, J.; Versteeg, G. F.; Brilman, D. W. F.; Hogendoorn, J.
A. Kinetic Study of CO2 with Various Amino Acid Salts in Aqueous
Solution. Chem. Eng. Sci. 2009, 64, 59.
(8) Portugal, A. F.; Derks, P. W. J.; Versteeg, G. F.; Magalhaes, F. D.;
Mendes, A. Characterization of Potassium Glycinate for Carbon
Dioxide Absorption Purposes. Chem. Eng. Sci. 2007, 62, 6534.
(9) Portugal, A. F.; Magalhaes, F. D.; Mendes, A. Carbon Dioxide
Absorption Kinetics in Potassium Threonate. Chem. Eng. Sci. 2008, 63,
3493.
(10) Simons, K.; Brilman, D. W. F.; Mengers, H.; Nijmeijer, K.;
Wessling, M. Kinetics of CO2 Absorption in Aqueous Sarcosine Salt
Solutions: Influence of Concentration, Temperature, and CO2
Loading. Ind. Eng. Chem. Res. 2010, 49, 9693.
(11) Paul, S.; Thomsen, K. Kinetics of Absorption of Carbon Dioxide
into Aqueous Potassium Salt of Proline. Int. J. Greenh. Gas Con. 2012,
8, 169.
(12) Majchrowicz, M. E.; Brilman, D. W. F. Solubility of CO2 in
Aqueous Potassium L-prolinate SolutionsAbsorber Conditions.
Chem. Eng. Sci. 2012, 72, 35.
(13) Caplow, M. Kinetics of Carbamate Formation and Breakdown. J.
Am. Chem. Soc. 1968, 90, 6796.
(14) Danckwerts, P. V. The Reaction of CO2 with Ethanolamines.
Chem. Eng. Sci. 1979, 34, 443.
(15) 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.
(16) Versteeg, G. F.; van Dijck, L. A. J.; van Swaaij, W. P. M. On the
Kinetics between CO2 and Alkanolamines both in Aqueous and Nonaqueous Solutions. An overview. Chem. Eng. Commun. 1996, 144, 113.
(17) Pohorecki, R.; Moniuk, W. Kinetics of the Reaction Between
Carbon Dioxide and Hydroxyl Ion in Aqueous Electrolyte Solutions.
Chem. Eng. Sci. 1988, 43, 1677.
(18) Danckwerts, P. V. Gas-Liquid Reactions. McGraw-Hill Book
Company: New York, 1970.
(19) Derks, P. W. J.; Kleingeld, T.; van Aken, C.; Hogendoorn, J. A.;
Versteeg, G. F. Kinetics of Absorption of Carbon Dioxide in Aqueous
Piperazine Solutions. Chem. Eng. Sci. 2006, 61, 6837.
(20) Hikita, H.; Asai, S. Gas Absorption with a Two Step Chemical
Reaction. Chem. Eng. J. 1976, 11, 123.
(21) Higbie, R. The Rate of Absorption of a Pure Gas into a Still
Liquid during a Short Time of Exposure. Trans. Am. Inst. Chem. Eng.
1935, 31, 365.

ASSOCIATED CONTENT

S Supporting Information
*

The density and dynamic viscosity of potassium, sodium, and


lithium salts of L-proline were measured over the temperature
range of 284313 K and concentrations of 0.1 M to the
maximum concentration (as given by Majchrowicz et al.33).
The solubility of N2O in aqueous L-prolinate solutions was
measured at the same range of temperatures and solutions
concentrations; showing typical salting out behavior. The
physical solubility of CO2 can then be calculated based on the
solubility of CO2 in water28 and the Schumpe model.27 The
results of this study are presented in Figures S1 and S2 and in
Tables S1 to S5. Tables S6 to S9 report the results of the
kinetic study for the reaction of CO2 with MEA and potassium
salt of L-proline. This material is available free of charge via the
Internet at http://pubs.acs.org.

R = universal gas constant (8.1343) [Jmol1K1]


T = temperature [K]
= loading of CO2 [mol CO2 per mol of AAS]
AAS = amino acid salt

AUTHOR INFORMATION

Corresponding Author

*E-mail: memajchrowicz@gmail.com.
Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
This research is part of the CATO Program, the Dutch national
research program on CO2 Capture and Storage. CATO is
nancially supported by the Dutch Ministry of Economic
Aairs (EZ) and the consortium partners (www.co2-cato.nl).

NOTATION
A = gasliquid interfacial area [m2]
B = base (H2O, OH, AAS)
Di = diusion coecient of component i [m2s1]
E = enhancement factor [-]
E = innite enhancement factor [-]
Ha = Hatta number [-]
hi = interaction parameter in Schumpe model for ions [dm3
mol1]
hg = interaction parameter in Schumpe model for gases
[dm3mol1]
JCO2 = CO2 absorption mole ux [molm2s1]
k1 = zwitterion mechanism rate constant [s1]
k2 = second order rate constant [dm3mol1s]
kapp = apparent rate constant [s1]
kAAS = zwitterion mechanism deprotonation rate constant for
AAS [dm3mol1s1]
kH2O = zwitterion mechanism deprotonation rate constant for
H2O [dm3mol1s1]
kL = liquid phase mass transfer coecient [ms1]
kOH = zwitterion mechanism deprotonation rate constant
for OH [dm3mol1s1]
kov = overall rate constant [s1]
m = physical distribution coecient [-]
n = partial order of the reaction in amino acid salt [-]
11466

dx.doi.org/10.1021/ie501083v | Ind. Eng. Chem. Res. 2014, 53, 1146011467

Industrial & Engineering Chemistry Research

Article

(22) Park, S. W.; Choi, B. S.; Oh, K. J.; Lee, J. W. Absorption of


Carbon Dioxide into Aqueous PAA Solution Containing Diethanolamine. J. Chin. Inst. Chem. Eng. 2007, 38, 461.
(23) Vucak, M.; Peric, J.; Z mikic, A.; Pons, M. N. A Study of Carbon
Dioxide Absorption into Aqueous Monoethanolamine Solution
Containing Calcium Nitrate in the GasLiquid Reactive Precipitation
of Calcium Carbonate. Chem. Eng. J. 2002, 28, 171.
(24) DeCoursey, W. J. Absorption with Chemical Reaction:
Development of a New Relation for the Danckwerts Model. Chem.
Eng. Sci. 1974, 29, 1867.
(25) Van Swaaij, W. P. M.; Versteeg, G. F. Mass Transfer
Accompanied with Complex Reversible Chemical Reaction in Gas
Liquid Systems: An Overview. Chem. Eng. Sci. 1992, 47, 3181.
(26) Laddha, S. S.; Diaz, J. M.; Danckwerts, P. V. The N2O Analogy:
The Solubilities of CO2 and N2O in Aqueous Solutions of Organic
Compounds. Chem. Eng. Sci. 1981, 36, 228.
(27) Weisenberger, S.; Schumpe, A. Estimation of Gas Solubilities in
Salt Solutions at Temperatures from 273 to 363 K. AIChE J. 1996, 42,
298.
(28) Versteeg, G. F.; van Swaaij, W. P. Solubility and Diffusivity of
Acid Gases (Carbon Dioxide, Nitrous Oxide) in Aqueous Alkanolamine Solutions. J. Chem. Eng. Data 1988, 33, 29.
(29) Gubbins, K. E.; Bhatia, K. K.; Walker, R. D. Diffusion of Gases
in Electrolytic Solutions. AIChE J. 1966, 12, 548.
(30) Van Holst, J.; Kersten, S. R. A.; Hogendoorn, K. J.
Physicochemical Properties of Several Aqueous Amino Acid Salts. J.
Chem. Eng. Data 2008, 53, 1286.
(31) Popiel, C. O.; Wojtkowiak, J. Simple Formulas for
Thermophysical Properties of Liquid Water for Heat Transfer
Calculations (from 0 to 150 C). HTrEng. 1998, 19, 87.
(32) Hamborg, E. S.; van Swaaij, W. P. M.; Versteeg, G. F.
Diffusivities in Aqueous Solutions of the Potassium Salt of Amino
Acids. J. Chem. Eng. Data 2008, 53, 1141.
(33) Majchrowicz, M. E.; Brilman, D. W. F.; Groeneveld, M. J.
Precipitation Regime for Selected Amino Acid Salts for CO2 Capture
from Flue Gases. Energy Procedia 2009, 1, 979.
(34) Laddha, S. S.; Danckwerts, P. V. Reaction of CO2 with
Ethanolamines: Kinetics from Gas-Absorption. Chem. Eng. Sci. 1981,
36, 479.
(35) Versteeg, G. F.; Oyevaar, M. H. The Reaction between CO2 and
Diethanolamine at 298 K. Chem. Eng. Sci. 1989, 44, 1264.
(36) Haubrock, J.; Hogendoorn, J. A.; Versteeg, G. F. The
Applicability of Activities in Kinetic Expressions: A More Fundamental
Approach to Represent the Kinetics of the System CO2-OHSalt in
Terms of Activities. Chem. Eng. Sci. 2007, 62, 5753.
(37) Crooks, J. E.; Donnellan, J. P. Kinetics and Mechanism of the
Reaction between Carbon Dioxide and Amines in Aqueous Solution. J.
Chem. Soc., Perkin Trans. 2 1989, 4, 331.
(38) Perry, R. H.; Green, D. W. Perrys Chemical Engineers Handbook,
6th ed.; McGraw-Hill: New York, 1997.

11467

dx.doi.org/10.1021/ie501083v | Ind. Eng. Chem. Res. 2014, 53, 1146011467

You might also like