You are on page 1of 20

molecules

Article
Study of Catalytic CO2 Absorption and Desorption
with Tertiary Amine DEEA and 1DMA-2P with the
Aid of Solid Acid and Solid Alkaline Chemicals
Huancong Shi 1,2, * , Min Huang 1 , Qiming Wu 1 , Linna Zheng 1 , Lifeng Cui 1, *,
Shuping Zhang 1, * and Paitoon Tontiwachwuthikul 2, *
1 Department of Environmental Science and Engineering, College of Science, University of Shanghai for
Science and Technology, Shanghai 200093, China; 172681815@st.usst.edu.cn (M.H.);
19921260976@163.com (Q.W.); linna1234576@163.com (L.Z.)
2 Clean Energy Technology Research Innovation (CETRI), Faculty of Engineering and Applied Science,
University of Regina, 3737 Wascana Parkway, Regina, SK S4S 0A2, Canada
* Correspondence: hcshi@usst.edu.cn (H.S.); lifengcui@gmail.com (L.C.);
zhang_lucy9999@vip.126.com (S.Z.); paitoon@uregina.ca (P.T.)

Academic Editor: Carlo Nervi 



Received: 21 January 2019; Accepted: 8 March 2019; Published: 13 March 2019

Abstract: Studies of catalytic CO2 absorption and desorption were completed in two well-performed
tertiary amines: diethylmonoethanolamine (DEEA) and 1-dimethylamino-2-propanol (1DMA-2P),
with the aid of CaCO3 and MgCO3 in the absorption process, and with the aid of γ-Al2 O3 and
H-ZSM-5 in the desorption process. The batch process was used for CO2 absorption with solid
alkalis, and the recirculation process was used for CO2 desorption with solid acid catalysts. The CO2
equilibrium solubility and pKa were also measured at 293 K with results comparable to the literature.
The catalytic tests discovered that the heterogeneous catalysis of tertiary amines on both absorption
and desorption sides were quite different from monoethanolamine (MEA) and diethanolamine
(DEA). These results were illustrative as a start-up to further study of the kinetics of heterogeneous
catalysis of CO2 to tertiary amines based on their special reaction schemes and base-catalyzed
hydration mechanism.

Keywords: CO2 absorption; tertiary amines; CO2 desorption; energy cost; equilibrium; solubility

1. Introduction
The global warming and sudden change of climates have driven scientists and engineers to
develop cost-effective processes for CO2 removal from coal-fired power plants [1]. The chemical
absorption of CO2 in the post-combustion CO2 capture process may be implemented on the commercial
scale [2]. This absorption process enables the CO2 removal with “energy efficient” amine solvents via
an absorption-desorption unit [1].
The development of attractive and novel amines has been a strong drive to meet these basic
requirements: high absorption rates, large cyclic capacity, and low regeneration energy [3,4]. Amine
solutions of monoethanolamine (MEA), diethanolamine (DEA) and methyldiethanolamine (MDEA)
have been widely used for CO2 removal, as benchmarks of primary, secondary, and tertiary amines [5].
MEA exhibits a higher reaction rate, but smaller cyclic capacity, higher energy costs for regeneration
and higher corrosion rates [4]. To overcome these limitations, MEA is usually blended with a
variety of tertiary amines in preparation for improved solvents called “MEA-R3 N blends” such
as MEA-MDEA [6,7], MEA-4-diethylamino-2butabol (DEAB) [6,7], MEA-diethylmonoethanolamine
(DEEA) [8] and MEA-1-dimethylamino-2-propanol (1DMA-2P), etc. [9]. Among these MEA-R3 N
blends, the concentration of MEA is usually 5.0 mol/L, but the concentration of tertiary amine is

Molecules 2019, 24, 1009; doi:10.3390/molecules24061009 www.mdpi.com/journal/molecules


Molecules 2019, 24, 1009 2 of 20

around 1.0–2.0 mol/L for MDEA [6,7], and 1.0–1.5 for amines such as DEEA (1.0 mol/L) [9], 1DMA-2P
(1.0 mol/L) [9] of DEAB (1.25 mol/L) [6,7] etc. In industry, the tertiary amines are usually prepared
at a concentration of 1–1.5 mol/L for absorption if blended with a concentrated amine such as
5.0 mol/L MEA.
Most tertiary amines fulfill two basic requirements: a large cyclic capacity with a relatively lower
regeneration energy compared with MEA and DEA [5]. Among these commercial tertiary amines,
MDEA is always used as benchmark. Recently, two good performance tertiary amines have drawn
research interest: DEEA and 1DMA-2P [1,5]. DEEA is studied for its relatively higher CO2 equilibrium
solubility than MDEA [10]. The pKa of it [11] is studied, and its mass transfer performance [12] is
also better than MDEA. 1DMA-2P is also investigated for its large CO2 cyclic capacity, much higher
absorption rates (kinetics) [13] and better mass transfer characteristics [14] than MDEA. It also has a
much lower CO2 absorption heat than MEA, DEA, PZ and MDEA [15]. Moreover, the equilibrium
solubility, the pKa [16] and ion speciation plots of 1DMA-2P-CO2 -H2 O are generated with 13 C NMR
methods [17].
Meanwhile, based on recent studies, the effects of solid base chemicals to MEA (K/MgO, CaCO3
and MgCO3 ) [18,19] and DEA (CaCO3 and MgCO3 ) [20] on CO2 absorption have been verified. They
accelerate the CO2 absorption rates. However, the effects of both solid bases onto a tertiary amine
have rarely been discussed either. The effects of solid acid catalysts (γ-Al2 O3 , H-ZSM-5, TiO(OH)2 ,
and transition metal oxides V2 O5 , MoO3 , WO3 , TiO2 , and Cr2 O3 ) to MEA [9,21–25] and DEA [26] have
also been studied and proven to be effective in the reduction of heat duty. Some studies have been
completed to investigate the effects of solid acid catalysts (H-ZSM-5, MCM-41 and SO4 2− /ZrO2 ) with
blended amine solvents of 5 + 1.0 mol/L MEA-DEEA and MEA-1DMA-2P [9]. However, the effects of
the heterogeneous catalysis toward CO2 -R3 N alone require detailed investigation.
The catalytic effects were verified of CaCO3 and MgCO3 toward MEA (19) and DEA (20) on CO2
absorption, and also (γ-Al2 O3 and H-ZSM-5) to MEA [9,21–23] and DEA [26] on CO2 desorption.
These catalytic effects toward tertiary amines were the purpose of this study. Since the reaction
schemes and the mechanism of CO2 absorption with tertiary amines are quite different from primary
amine (MEA) or secondary amine (DEA), these differences make the effects of heterogeneous catalysis
worthy of deep investigation with solid bases and solid acids as a start-up. For this study, the CO2
absorption with 1.0–1.5 mol/L DEEA and 1DMA-2P solvents were investigated with the aid of a solid
base (CaCO3 and MgCO3 ), and CO2 desorption 1.0–2.0 mol/L DEEA and 1DMA-2P solvents for solid
acids (γ-Al2 O3 and H-ZSM-5). The effects of heterogeneous catalysis on both sides of absorption and
desorption were studied and compared with MEA and DEA.

2. Theory

2.1. Reaction Scheme, and Suitable Mechanisms of CO2 -R3 N Interaction


The reaction scheme of the CO2 reaction with tertiary amine (R1 R2 R3 N) is presented below
with Equations (1)–(6) [3]. Equation (1) is the major reaction being emphasized. Different from
primary/secondary amines (R1 NH2 /R1 R2 NH), the major anion is bicarbonate (HCO3 − ) instead of
carbamate (R1 R2 N-COO− ).

CO2 + R1 R2 R3 N + H2 O ↔ R1 R2 R3 NH+ + HCO3 − (1)

R1 R2 R3 NH+ ↔ R1 R2 R3 N + H+ (2)

CO2 + H2 O ↔ H2 CO3 ↔ H+ + HCO3 − (3)

CO2 + OH− ↔ HCO3 − (4)

HCO3 − ↔ H+ + CO3 2− (5)

H2 O ↔ H+ + OH− (6)
Molecules 2019, 24, 1009 3 of 20

Based on a recent review of kinetics [27], the Zwitterion mechanism [28] and Termolecular
mechanism [29] are suitable for CO2 reactions with primary and secondary amines. The mechanism
of CO2 reaction with tertiary amines was proposed by Donaldson and Nguyen and is termed the
“base-catalyzed hydration mechanism” [30]. The tertiary amine (R1 R2 R3 N) does not react directly with
CO2 , but rather acts as a base that catalyzes the hydration of CO2 [30] based on Equation (1).
The rate equation was written as Equation (7), with the rate constant of tertiary amine (R3 N) of
kR3 N : [31] from Equation (7), the bigger amine concentration results in higher absorption rates.

rCO2 = kR3 N [R3 N][CO2 ]. (7)

Moreover, there are three balance equations in the R3 N-CO2 -H2 O systems, as presented in
Equations (8)–(10) [3].
Mass balance of R3 N:

[R1 R2 R3 N]0 = [R1 R2 R3 NH+ ] + [R1 R2 R3 N] (8)

Mass balance of Carbon, where α is CO2 loading:

α × [R1 R2 R3 N]0 = [CO2(aq) ] + [HCO3 − ] + [CO3 2− ] (9)

Charge balance:

[R1 R2 R3 NH+ ] + [H+ ] = [OH− ] + [HCO3 − ] + 2[CO3 2− ] (10)

2.2. Role of Solid Alkalis Chemicals for Absorption


The catalytic effects of both solid alkalis to MEA and DEA were already studied [19,20]. However,
the role was slightly different with tertiary amines. Since the reaction mechanism of CO2 -R3 N was
“base-catalyzed hydration” [30], reactions (1)–(6) could be facilitated with either a liquid base [OH− ]
or solid alkalis/Lewis base [19] due to the acidic chemical nature of CO2 . Therefore, the solid alkalis
(CaCO3 and MgCO3 ) could enhance the hydration of CO2 , and proton transfer of [H2 CO3 ] to water or
R3 N, as an aid to the liquid base of [OH− ].

Base
CO2 + H2 O ↔ H+ + HCO3 − (11)

Base
CO2 + R1 R2 R3 N + H2 O ↔ H+ + HCO3 − (12)

After detailed investigations, the role of solid alkalis was “Lewis base” and “proton acceptor” [19]
to facilitate proton transfer of H2 CO3 , with reactions below:

H2 CO3 + Catalyst ↔ HCO3 − + Cat-[H+ ] (13)

Cat-[H+ ] + R3 N ↔ Catalyst + [R3 NH+ ] (14)

The solid catalyst (CaCO3 /MgCO3 ) accepted the protons [H+ ] from H2 CO3 with its long pair
of electrons on the “O atom” via (13). After that, the proton was transferred from the solid catalyst
to the tertiary amine (R3 N) via (14), because the tertiary amine is a stronger base than Lewis base
(CaCO3 /MgCO3 ). The overall reaction was (3) + (13) + (14) = (1), and the solid catalyst was involved
in the reaction but did not change the reactant or products. From Equations (13) and (14), with an
increased amount of solid base, the reaction rate of (13) would increase but the rate of (14) would
decrease. With an increased mass of catalysts, the protons [H+ ] are easier to transfer onto a solid
surface, but harder to release back to R3 N. Therefore, there is an optimized amount of solid catalyst for
CO2 absorption, after that, the rates slightly decreased.
Molecules 2019, 24, 1009 4 of 20

2.3. Role of Lewis Acid and BrØnsted Acid for CO2 Desorption
The role of Lewis acid (γ-Al2 O3 ) and BrØnsted acid for CO2 desorption with MEA and DEA has
already been discussed repeatedly [9,21,26].
However, the role of both acids needs to be discussed for CO2 desorption with tertiary amines
because the reactions were different, and no carbamate was involved. After the study of reaction
schemes, the effect and mechanism of both solid acids were discussed as follows:
γ-Al2 O3 as catalyst:
Al2 O3 + 2OH− ↔ 2 AlO2 − + H2 O (15)

R1 R2 R3 NH+ + AlO2 − ↔ R1 R2 R3 N + HAlO2 (16)

HAlO2 + HCO3 − ↔ AlO2 − + H2 O + CO2 ↑ (17)

H-ZSM-5 as catalyst:

H-ZSM-5 + HCO3 − ↔ (ZSM-5)− + H2 O + CO2 ↑ (18)

R1 R2 R3 NH+ + (ZSM-5)− ↔ R1 R2 R3 N + H-ZSM-5 (19)

Both solid acids could facilitate the CO2 production rates and reduce the energy costs. In short,
the Al2 O3 is Amphoteric Oxide, and it was converted into AlO2 − in basic solutions. From the published
energy diagrams [6], Al2 O3 (AlO2 − ) can speed up the proton transfer from R1 R2 R3 NH+ to HCO3 −
and CO2 generation under heat. H-ZSM-5 is the proton donor, which directly provides protons to the
solvent and facilitates CO2 generation.
Based on the reaction schemes above, both solid acids involve two steps, namely “accept proton”
and “donate/transfer proton”. Since γ-Al2 O3 contains no proton, it has to “accept proton” first and
“transfer proton” later in the desorption process. The H-ZSM-5 contains proton itself, and it intends
to “donate proton” first and then “accept proton” from R1 R2 R3 NH+ . The mechanisms are similar for
both solid catalysts but the order of “accept proton” and “donate/transfer proton” is opposite.

3. Materials and Experimental Methods

3.1. Chemicals
The solid chemicals were purchased from Huishan Chemical Ltd; they were CaCO3 and MgCO3 .
The CO2 gas and the liquid chemicals DEEA and 1-DMA2P were purchased from Tansoole Chemical
Ltd. (Shanghai, China). HCl and methyl orange were commercially obtained from Guoyao Chemical
Ltd. (Shanghai, China). The chemical structures and full name of DEEA and 1DMA2P were presented
elsewhere [5].

3.2. pKa Analysis


The titration technique was adopted to determine the amine dissociation constant (Ka) with
standard 1 mol/L HCl [32–35]. This is a simplified pH method to test the pKa of different amines
under different temperatures. For an aqueous amine solution, the Equation (20) below showed the
deprotonation reaction of AmineH+ /Amine as a conjugated pair of acid-bases.

k
AmineH+ ↔a Amine + H+ (20)

Based on a detailed pKa study of tertiary amines [11] the pH meter measured the activity {R3 NH+ }
of amine solvents instead of its real concentration [R3 NH+ ]. The correlation is {R3 NH+ } = [R3 NH+ ]
γBH+ [11]. This study assumed that this diluted amine solvent (<0.10 mol/L) is the ideal solution
(when the concentration is very low and the activity coefficient γBH+ equals to 1 {BH+ } ≈ [BH+ ]) [11].
Then, Equations (21) and (22) below were used to calculate the Ka.
Molecules 2019, 24, 1009 5 of 20

[Amine] H+
 
Ka,Amine =  (21)
AmineH+


pKa = − log(Ka,Amine ) = − log [Amine] H+ / AmineH+


   
(22)

The pH meter was used to measure the concentration of H+ in the solution [32–35]. As presented
in Equation (23), the disappearance of H+ during the titration process resulted from its reaction with
Amine to generate AmineH+ , and reaction (20) is the dominant in aqueous solution. A mass balance of
protons as shown in Equation (23) was adopted to calculate the concentration of AmineH+ , and the
amine balance equation as shown in Equation (24) was adopted to calculate free amine.

nHCl − H+ Vtotal = AmineH+ Vtotal


   
(23)

[Amine] + AmineH+ Vtotal = n0,Amine


 
(24)

In Equations (23) and (24) above, nHCl is the number of moles of HCl added during the titration
process, Vtotal is the total liquid volume after the titration process, and n0,Amine is the moles of free
amines as a start, which can be determined by titration with 1.0 mol/L HCl until the indicator of
methyl orange turns pink.
For the experiment, the Ka of amine was determined based on the procedure described [32].
Briefly speaking, 100 mL of 0.10 mol/L amine solution was carefully prepared and titrated with 10 mL
of 1.0 mol/L HCl standard solution at 298 K until the end point was observed. During the titration
process, the pH meter was placed in the solution to record pH value with the addition of 1 mL HCl
each time. A table of pH value and amount of HCl was generated. Equations (21)–(24), were used to
determine the concentration of [AmineH+ ], [Amine] and then calculate the dissociation constant (Ka).
The dataset was only recorded at pH > 9, because the results would be inaccurate if pH < 9, where the
generation of [H+ ] or [OH− ] from water is not negligible.

3.3. CO2 Absorption Process with Absorption Profiles


A set of stirred-cell reactors were built in the lab, with the flow chart exhibited in Figure S1.
The process was similar to that of other studies, [17,20] and the diameter of the reactor is 11.0 cm
(a constant interfacial area of 95.0 cm2 ). The solid alkalis (CaCO3 and MgCO3 ) were pelletized at
2–3 mm and wrapped into small balls with a diameter of 2.5 cm, each ball contained about 2.5 g,
similarly [20].
The CO2 absorption process was similar to that of other works [17,20]. Three-hundred milliliters
of amine solvent was prepared at a concentration of 1.0–1.5 mol/L. For 1DMA-2P, it started to crystalize
at 2.0 mol/L at 293 K, then 2.0 mol/L was not tested for absorption. The CO2 gas flow was introduced
into the reactor at a rate of 1.5 L/min. The PCO2 was 101.3kPa with 100% purity. The operation
temperature was maintained at 20 ◦ C by the cooled water bath. The process was connected to the air
with the pressure of 1 atm. Some vials were prepared and 2 mL samples were pipetted every 3–5 min
into each of them. The titration technique was adopted to test the CO2 loadings and the results were
recorded within 3−5 min [17]. A Chittick apparatus with an AAD of 2.5% was adopted to conduct the
CO2 -loading tests of the samples [36]. In order to ensure repeatability, these tests were performed at
least twice.
We already verified the catalytic effect of CaCO3 with CO2 absorption with 1.0 mol/L MEA in
another study [19]. The results exhibited that the order of catalytic effects was: 5 g CaCO3 in gas-liquid
interface >5 g, CaCO3 in bulk of liquid >0 g, CaCO3 ≈ intert stainless steel. From the BET tests
of CaCO3 and MgCO3 , the surface areas were 0.428 m2 /g for CaCO3 and 9.498 m2 /g for MgCO3 .
The pore diameters were 31.3 nm and 4.31 nm, which facilitated the external mass transfer of amine
molecules onto a solid surface. The inert material with large surface area might have a significant
effects of mass transfer causing the increase of CO2 absorption. However, it could not replace the role
Molecules 2019, 24, 1009 6 of 20

of “Lewis base“ or “proton acceptor“ as solid alkaline catalysts to enhance the CO2 absorption with
tertiary amine.
After the absorption profiles were plotted with (α, time), the initial absorption rates (Iabs_rate ,
mol CO2 /min) [34] are determined as the slope of the linear regression of absorption profiles was at
data range of CO2 loadings of 0.0–0.20 mol/mol, in Equation (25):


Iabs_rate = C × V (25)
dt
where “C” is the Concentration and “V” is the Volume, and α is the CO2 loading.
The initial absorption rates were adopted in this study to compare the CO2 absorption
performance of different cases of catalysts. These data were generated at a consistent level for different
catalytic and non-catalytic cases. The results were inadequate for kinetic studies for now, but it was
adequate to verify the catalytic effect as a start up.

3.4. CO2 Desorption Tests with Heat Duty Calculation


An open recirculation-process (Figure S2) vessel equipped with an electrometer [9,26,37] was
adopted to conduct the CO2 desorption tests under atmosphere to extract DEEA and 1DMA-2P
solvents at 1.0 mol/L, 1.5 mol/L and 2.0 mol/L, and two types of solid acid catalyst were used
as γ-Al2 O3 and H-ZSM-5 representing Lewis acids and Brønsted acids [9,26]. The acidic catalytic
conditions were 5.0 g, 7.5 g and 10.0 g, This CO2 desorption process was similar to that of others in the
literature [9,26]. In this study, 250 mL of the amine solvent was put into the flask with a volume of
500 mL. The CO2 loading was over 0.80 mol/mol in preparation for desorption, with CO2 introduced
into amine solvents beforehand. Small balls of various catalysts were placed into the solvents as well.
The experimental procedures were similar to those in our previous study [26]. The process was
stirred and heated to 363 K. Based on the analysis of the CO2 loading of samples at 0−4 h, the catalytic
effects on CO2 desorption were evaluated. A vial into which the samples were pipetted was then
cooled down in a cooled water bath so as to maintain the CO2 loading. Similar to absorption, the CO2
loading was tested immediately after sample collection by titration [26]. A Chittick apparatus [36] was
adopted to conduct the CO2 -loading tests of each sample and they were performed at least twice so as
to ensure repeatability. The average CO2 loading was then plotted for each run.
As part of the pivotal desorption parameter [9,26,33], the heat duties were calculated from
CO2 production with Equations (26) and (27) below [9,21,26]. The nCO2 (mol) is the amount of
desorbed CO2 , α (mol of CO2 per mole of amine) is the CO2 loading, and C (mol/L) and V (L) are the
concentration and volume of amine solvent. This method of calculating heat duties was similar to that
in other studies [9,21,26].

heat input/time E kJ of electricity/h


H= = (26)
amount of CO2 /time nCO2 mol of CO2 /h

nCO2 = (αrich − αlean )C × V (27)

4. Results and Discussions

4.1. The Critical Point of CO2 Absorption Curve of DEEA and 1-DMA-2P at 293 K, Affected by Equilibrium
Solubility
The CO2 equilibrium solubility at different temperatures and pressures was one of the key
parameters for screen solvent [15]. Although different solid base chemicals could accelerate the CO2
absorption rates, they could hardly shift the CO2 equilibrium solubility under the VLE model, which
was only determined by temperature [10].
There are two common methods to generate the VLE model of tertiary amines, (1) the predictive
model by simulation using MDEA as benchmark [38,39], (2) experimental studies of CO2 solubility
Molecules 2019, 24, 1009 7 of 20

with absorption [5,10,15]. The equilibrium solubility of DEEA and 1DMA-2P was reported based on
long-term vapor liquid equilibrium experiments at 298–313 K [5]. The accurate equilibrium solubility
was relatively high
Molecules 2019, duePEER
24, x FOR to the experimental procedures. Luo et al. completed the experiments
REVIEW 7 of 22 and
the modeling of data of DEEA-CO2 -H2 O at 1.0 to 4.0 mol/L under vapor-liquid equilibrium, from
long-term
293–353 K [10].vapor
Theyliquid
used equilibrium
0.15 L/minexperiments
mixed gas of at 298–313
CO2 /NK [5]. The accurate equilibrium solubility
2 , with CO2 partial pressures from 6.2 kPa
was relatively high due to the experimental procedures. Luo et al. completed the experiments and
to 100.8 kPa and maintained 8–10 h to reach the thermodynamic equilibrium conditions. By the
the modeling of data of DEEA-CO2-H2O at 1.0 to 4.0 mol/L under vapor-liquid equilibrium, from 293–
long-term tests, the equilibrium solubility of DEEA was 0.971 mol/mol at 293 K with PCO2 of 100.8 kPa.
353 K [10]. They used 0.15 L/min mixed gas of CO2/N2, with CO2 partial pressures from 6.2 kPa to
Similarly, Liu et al. completed the modeling of CO2 equilibriumequilibrium
100.8 kPa and maintained 8–10 h to reach the thermodynamic
solubility of 1DMA-2P solution, with
conditions. By the long-
CO2 partial
term tests, the equilibrium solubility of DEEA was 0.971 mol/mol at 293 K withKP[15].
pressures from 8 kPa to 101.3 kPa at 1, 2, and 5 mol/L, 298–333 For 1.0 mol/L
CO2 of 100.8 kPa.
1DMA-2P,
Similarly, Liu et al. completed the modeling of CO2 equilibrium solubility of 1DMA-2P solution, with[15].
the solubility was reported as 1.02 mol/mol at 101.3 kPa at 298 K with 8 h operation
In
COthis study,
2 partial we determined
pressures from 8 kPa theto“critical
101.3 kPa point
at 1,of2,COand2 absorption curves”
5 mol/L, 298–333 based
K [15]. Foron1.0the slope of
mol/L
1DMA-2P, the
CO2 absorption solubility
curves of 1.0 was reported
mol/L amineas 1.02 mol/mol
without at 101.3 It
catalysts. kPa at 298
was K with
affected by8 the
h operation [15].
CO2 solubility or
In this plot”
“Ion speciation study,of weRdetermined
3 N-CO 2 -H the
2 O “critical
systems point
under of CO
the absorption
2Vapor-liquid curves” based
equilibrium on the
model.slope
Fromof the
CO2 absorption
CO2 absorption curves
curves, of 1.0 mol/L
different stages amine
of thewithout
reactioncatalysts. It was affected
were contained, (1) by
COthe CO2 solubility or
2 + R3 N + H2 O around
“Ion speciation plot” of R3N-CO2-H2O systems under the Vapor-liquid equilibrium model. From the
0–0.85 mol/mol, and (2) CO2 + H2 O above 0.85 mol/mol when free R3 N was exhausted. The slope of
CO2 absorption curves, different stages of the reaction were contained, (1) CO2 + R3N + H2O around
absorption curves turned very flat after this critical point, indicating that all the amines were converted
0–0.85 +mol/mol, and (2) CO2 + H2O above 0.85 mol/mol when free R3N was exhausted. The slope of
to amineH , and CO only reacted with water afterward. The critical point was determined by the
absorption curves2 turned very flat after this critical point, indicating that all the amines were
graphic
converted to based
method amineH on the CO
+, and cross of slopes at different stages (Support Information).
2 only reacted with water afterward. The critical point was determined
Therefore,
by the graphic themethod
“critical point
based onof theCO 2 absorption
cross of slopes atcurves”
different was calculated
stages as about 0.87 mol/mol
(Support Information).
for DEEA,Therefore,
and 0.81the mol/mol for 1DMA-2P
“critical point at 1.0 mol/L
of CO2 absorption curves”and was293 K hereaswas
calculated based
about on the graphic
0.87 mol/mol for
DEEA, and 0.81 mol/mol for 1DMA-2P at 1.0 mol/L and 293 K
method. The CO2 solubility of DEEA and 1DMA-2P under Vapor-Liquid Equilibrium (VLE) model here was based on the graphic method.
The COin
was plotted 2 solubility
Figure 1of at DEEA
298–313 and K 1DMA-2P
[5]. Our data under Vapor-Liquid
were added in FigureEquilibrium
1, but(VLE)
thesemodel was not
data were
plotted in Figure 1 at 298–313 K [5]. Our data were added in Figure 1,
“CO2 equilibrium solubility”. It was affected by the “CO2 equilibrium solubility” and “Ion speciationbut these data were not “CO 2

equilibrium solubility”. It was affected by the “CO2 equilibrium solubility” and “Ion speciation
plots”. From the literature value, CO2 equilibrium solubility of CO2 is 0.839 mol/mol for DEEA,
plots”. From the literature value, CO2 equilibrium solubility of CO2 is 0.839 mol/mol for DEEA, and
and 0.789 mol/mol for 1DMA-2P at 298 K [5]. The trend was consistent from Figure 1 at 298–313 K:
0.789 mol/mol for 1DMA-2P at 298 K [5]. The trend was consistent from Figure 1 at 298–313 K: with
with an increaseofoftemperature,
an increase temperature, thethe solubility
solubility of CO of2 CO was slightly
was2slightly decreasing
decreasing [5,10]. [5,10].

0.90
(mol CO2 / mol amine )

0.85
DEEA
CO2 loading

0.80 1DMA-2P

0.75

0.70

0.65
288 293 298 303 308 313 318
T (K)
Figure 1. The
Figure 1. critical pointpoint
The critical of COof2 absorption curves
CO2 absorption at 1 atm
curves at 1and
atm293
andK,293
andK,CO 2 equilibrium
and solubility
CO2 equilibrium
of DEEA and 1DMA-2P
solubility of DEEA andat 1DMA-2P
pressure of 1 atm and
at pressure of 298–313
1 atm andK298–313
[5]. K [5].

4.2. The
4.2.pKa
The of
pKaDEEA andand
of DEEA 1-DMA-2P
1-DMA-2Patat293
293KK
The pKa is also
The pKa an important
is also an important parameter
parameter for
fortertiary
tertiaryamines,
amines, which canbe
which can beused
usedfor
forthethe selection of
selection
amineofsolutions for both
amine solutions forCOboth
2 removal
CO 2 and
removal the
and study
the studyof the
of reaction
the reactionkinetic
kinetic mechanism
mechanism [15].
[15]. Based on
Based
on the base-catalyzed mechanism, tertiary amines (R N) do not directly react
the base-catalyzed mechanism, tertiary amines (R3 N) do not directly react with CO2 , but absorbed
3 with CO 2 , but absorbed
protons
protons fromfromH2 CO H2CO 3. The simplified pH method for the detection of pKa is quite similar to that of
3 . The simplified pH method for the detection of pKa is quite similar to that of
other studies [32–35].
other studies [32–35]. It excluded It excluded the
the dataofofpH
data pH<< 9.0.
9.0. This
Thiswas
wasbecause
because the conjugated
the conjugated acid/base of of
acid/base
[Amine]/[AmineH +] did not exist under acidic conditions (pH < 7.0). Moreover, a pH value between
[Amine]/[AmineH+ ] did not exist under acidic conditions (pH < 7.0). Moreover, a pH value between
7–9 was not selected either, for the calculation of pKa in Equation (20) was based on the assumption
7–9 was not selected either, for the calculation of pKa in Equation (22) was based on the assumption
that the [H+] released into the solution was 100% from [AmineH+] with neglect of proton release from
that the + ] released into the solution was 100% from [AmineH+ ] with neglect of proton release
H2O.[HMeanwhile, the [OH−] in the water solution was mainly from proton transfer from H2O to
from Amine,
H2 O. Meanwhile, the [OH − ] in the water solution was mainly from proton transfer from
and [OH−] released from H2O was also negligible. In the case of pH < 9, the [OH−] in the H2 O
to Amine, and [OH − ] released from H O was also negligible. In the case of pH < 9, the [OH − ] in the
solution was smaller than 10 mol/L, −5
2 which was not 100 times bigger than the [OH ] (10 mol/L)
− −7
Molecules 2019, 24, 1009 8 of 20

Molecules
solution was2019, 24, x FOR
smaller PEER
than −5 mol/L, which was not 100 times bigger than the [OH− ] (10−
10REVIEW 8 7ofmol/L)
22

dissociated by neutral H2 O. Then the dissociation of [OH− ] from H2 O was not negligible, and thus
dissociated by neutral H2O. Then the dissociation of [OH ] from H2O was not negligible, and thus
Equation (22)(20)
Equation hadhad errors.
errors.Finally,
Finally, the pKawas
the pKa wasmeasured
measuredandand grouped
grouped into Table
into Table 1 and 2Figure
1 and Figure for 2
for comparison.
comparison.

Table pKa
1. 1.
Table ofofinvestigates
pKa investigates amines at293
amines at 293KKand
and 1 atm.
1 atm.

Amine Predicted pKa Measured


ReferenceReference Measured This WorkThis
Amine Predicted pKa
Work
9.73 (298 K) [11]
9.73 (298 K) [11]
DEEA
DEEA 9.60 9.60
(298(298
K) [5]
K) [5] 9.82 (2939.82
K) (293 K)
9.82 (293 K) [11]
9.82 (293 K) [11]
9.67 (301 K) [17]
9.67 (301 K) [17]
1DMA-2P
1DMA-2P 9.20 9.20
(298(298
K) [5]
K) [5] 9.51 (2939.51
K) (293 K)
9.41 (298 K) [15]
9.41 (298 K) [15]

9.6
y = 2639x + 0.559
9.4 R² = 1

9.2

9.0
pKa

8.8

8.6 y = 2545x + 0.8498


R² = 0.9978
8.4
0.0029 0.003 0.0031 0.0032 0.0033 0.0034 0.0035
1/T

2. The pKa of of
1DMA-2P a . a The equation above is for the pKa at
Figure
Figure 2. The pKa 1DMA-2Pat at 293–333
293–333 KKand
and 1 atm
1 atm a. a The equation above is for the pKa at 298–
333 K
298–333 K and
and 11atm
atmfrom
fromliterature, [15]
literature, andand
[15] the the
equation below
equation is the is
below calibration curve ofcurve
the calibration 4 points
of 4atpoints
the at
rangeof
the range of293–333
293–333 K.
K.

Hence, Hence, the pKa


the pKa was measured
was measured as 9.82
as 9.82 for DEEA
for DEEA at K.
at 293 293This
K. This
valuevalue wassame
was the the same
as theasliterature
the
literature value of 9.82 at 293 K, [11] which reflected the accuracy. The pKa was measured
value of 9.82 at 293 K, [11] which reflected the accuracy. The pKa was measured as 9.51 for 1DMA-2P as 9.51 for at
1DMA-2P at 293 K, comparable to 9.41 at 298 K [15] from K2 correlation model for solubility study. It
293 K, comparable to 9.41 at 298 K [15] from13K2 correlation model for solubility study. It was measured
was measured as 9.67 at 301 K based on C NMR analysis. [17] Different methods might result in
as 9.67 at 301 K based on 13 C NMR analysis. [17] Different methods might result in slight deviations.
slight deviations. Recently, Liu et al developed a linear calibration of pKa of 1DMA-2P in Equation
Recently, Liu
(25) at et al developed
298–333 K [15]. Weacombined
linear calibration
these data ofwith
pKaour
of 1DMA-2P
own results inand
Equation
plotted(27) at 298–333
in Figure 2. Our K [15].
We combined
data (9.51, these
293 K) data with our
was outside the own
rangeresults and plotted
of that calibration in Figure
curve, 2. Our
but the data wasdata (9.51,with
consistent 293the K) was
outside
line. The new linear calibration was generated in Equation (26) to expand the pKa of 1DMA-2P atlinear
the range of that calibration curve, but the data was consistent with the line. The new
calibration
293-333was
K. generated in Equation (28) to expand the pKa of 1DMA-2P at 293-333 K.
2639
pKa = 2639+ 0.559; 298 − 333 K (26)
pKa = T + 0.559; 298 − 333 K (28)
T
2545
pKa = + 0.850 293 − 333 K (27)
T
2545
pKa = + 0.850; 293 − 333 K (29)
T
4.3. The CO2 Absorption Profiles with Initial Absorption Rates
4.3. The CO2 Absorption Profiles with Initial Absorption Rates
The CO2 absorption profiles of 1.0 mol/L and 1.5 mol/L DEEA and 1DMA-2P solvents were
The COin2 Figures
plotted absorption profiles
3–6, with of 1.0
the aid mol/L
of CaCO and
3 and 1.5 3mol/L
MgCO DEEAThe
, respectively. andoptimized
1DMA-2P solvents
mass of solidwere
plotted in Figures
alkalis 3–6, with
under various theconcentrations
amine aid of CaCO3was and MgCO3in
presented , respectively. The optimized
Table 2. The optimized mass
mass was of solid
based
alkalis under various amine concentrations was presented in Table 2. The optimized mass wasofbased
on the catalytic reactions (A1) and (A2), with explanation in Section 2.2. With an increased mass
on the catalytic reactions (13) and (14), with explanation in Section 2.2. With an increased mass of solid
catalysts, the initial absorption rates increased first, reached an optimum and then decreased after that.
Molecules 2019, 24, x FOR PEER REVIEW 9 of 22

Molecules 2019,
solid 24, 1009 the initial absorption rates increased first, reached an optimum and then decreased9 of 20
catalysts,
after that.

A Catalytic CO2 absorption of 1.0 mol/L


DEEA solvents with CaCO3
1.00

Loading: mol CO2 / mol amine


0.90
0.80
0.70
0.60
0.50
10g CaCO3
0.40
7.5g CaCO3
0.30
5g CaCO3
0.20
0g CaCO3
0.10
0.00
0 5 10 15 20 25 30 35 40 45 50
Time (min)
Catalytic CO2 absorption of 1.0 mol/L
B
DEEA solvents with MgCO3
1.00
0.90
Loading: mol CO2 / mol amine

0.80
0.70
0.60
0.50
10g MgCO3
0.40
7.5g MgCO3
0.30
5g MgCO3
0.20
0g MgCO3
0.10
0.00
0 5 10 15 20 25 30 35 40 45 50
Time (min)
Figure
Figure 3. Catalytic
3. Catalytic COCO 2 absorption curves of 1.0 mol/L DEEA solvents at 293 K and 1 atm. (A) CaCO3
2 absorption curves of 1.0 mol/L DEEA solvents at 293 K and 1 atm. (A) CaCO3
0–10 g;
Molecules (B) MgCO
0–10 g; (B) MgCO3 0–10 g. g.REVIEW
2019, 24, x 3 0–10
FOR PEER 10 of 22

Catalytic CO2 absorption of 1.5 mol/L B Catalytic CO2 absorption of 1.5 mol/L
A
DEEA solvents with CaCO3 DEEA solvents with MgCO3
1.00 1.00
Loading: mol CO2 / mol amine
Loading: mol CO2 / mol amine

0.90 0.90
0.80 0.80
0.70 0.70
0.60 0.60
0.50 0.50
10g CaCO3 10g MgCO3
0.40 0.40
7.5g CaCO3 7.5g MgCO3
0.30 0.30
5g CaCO3 5g MgCO3
0.20 0.20
0g CaCO3 0g MgCO3
0.10 0.10
0.00 0.00
0 5 10 15 20 25 30 35 40 45 50 55 60 65 0 5 10 15 20 25 30 35 40 45 50 55 60 65
Time (min) Time (min)

Catalytic
Figure 4.Figure CO2 absorption
4. Catalytic curves
CO2 absorption of of
curves 1.51.5mol/L
mol/L DEEA solvents
DEEA solvents at at
293293 K and
K and 1 atm.
1 atm. (A) 3CaCO3
(A) CaCO
0–10 g (B) MgCO
0–10 g (B) MgCO3 0–10 g. 3 0–10 g.

Table 2. The optimized mass of CaCO3 and MgCO3 for different amine solvents at 293 K and 1 atm.

Amine Solvents CaCO3 (g) MgCO3 (g)


1.0 M DEEA 7.5 10
1.5 M DEEA 7.5 5
1.0 M 1DMA-2P 7.5 10
1.5 M 1DMA-2P 10 5

The non-catalytic curves for DEEA were plotted in Figures 3 and 4. In this time period, amine
Molecules 2019, 24, 1009 10 of 20

Table 2. The optimized mass of CaCO3 and MgCO3 for different amine solvents at 293 K and 1 atm.

Amine Solvents CaCO3 (g) MgCO3 (g)


1.0 M DEEA 7.5 10
1.5 M DEEA 7.5 5
1.0 M 1DMA-2P 7.5 10
1.5 M 1DMA-2P 10 5

The non-catalytic curves for DEEA were plotted in Figures 3 and 4. In this time period, amine
absorption was recorded from fresh solvent to equilibrium solubility of 0.87 mol/mol, the rest of the
data was not displayed because CO2 was reacting with H2 O. It took 40 min for 1.0 mol/L and 45 min
for 1.5 mol/L. With the aid of solid alkalis CaCO3 , the time was reduced to 24 min (21 min less) for
1.0 mol/L and 30 min (15 min less) for 1.5 mol/L at optimized conditions. With the aid of MgCO3 ,
the time was reduced to 24 min (21 min less) for 1.0 mol/L and 45 min for 1.5 mol/L at optimized
conditions. The effect of MgCO3 was similar to that of CaCO3 at 1.0 mol/L, but not very helpful at
1.5 mol/L with 5 g. If bigger than 5 g MgCO3 , the absorption curves were worse than the non-catalytic
curves due to “agglomeration” [40]. Therefore, CaCO3 was a better solid for DEEA than MgCO3 .
The non-catalytic curves for 1DMA-2P were plotted in Figures 5 and 6. In this time period, amine
absorption was recorded from fresh solvent to equilibrium solubility of 0.81 mol/mol. It took 35 min
for 1.0 mol/L, 30 min for 1.5 mol/L. The bigger amine concentration was, the less time it would take,
due to faster absorption rates [5] and the smaller cyclic capacity (0.81 mol/mol). With the aid of CaCO3 ,
the time was reduced to 18 min for 1.0 mol/L, 27 min for 1.5 mol/L at optimized conditions. With
the aid of MgCO3 , the time was reduced to 17 min for 1.0 mol/L, 27 min for 1.5 mol/L at optimized
conditions. The effect of CaCO3 was comparable to that of MgCO3 . It was quite effective at 1.0 mol/L
when the time was2019,
Molecules reduced
24, x FORby 18REVIEW
PEER min, and the time was reduced by only 3 min at 1.5 mol/L.
11 of 22

Catalytic CO2 absorption for 1.0 mol/L B Catalytic CO2 absorption for 1.0
A 1-DMA-2P solvents with CaCO3
1.00 mol/L DMA-2P solvents with MgCO3
1.00
0.90 0.90
Loading: mol CO2 / mol amine

0.80
Loading: mol CO2 / mol amine

0.80
0.70 0.70
0.60 0.60
0.50 0.50
10g CaCO3 10g MgCO3
0.40 0.40
7.5g CaCO3 7.5g MgCO3
0.30 0.30
5g CaCO3 5g MgCO3
0.20 0.20
0g CaCO3 0g MgCO3
0.10 0.10
0.00 0.00
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50
Time (min) Time (min)

Figure 5. Catalytic
Figure 5. Catalytic CO2 absorption
CO2 absorption curvesofof1.0
curves 1.0 mol/L
mol/L 1DMA-2P solvents
1DMA-2P at 293 K at
solvents and293
1 atm. (A). 1 atm.
K and
CaCO3 0–10 g. (B). MgCO3 0–10 g.
(A). CaCO3 0–10 g. (B). MgCO3 0–10 g.

The optimized amount of solid base chemicals is presented in Table 2. The orders were different
under different amine Catalytic CO2 absorption
it was 7.5ofg1.5 mol/L
g > 51DMA-2P
A concentrations. For DEEA, > 10 g > 0 g for 1.0 mol/L and
7.5 g ≈ 5 g > 0 g > 101.00 solvents with CaCO
g for 1.5 mol/L for CaCO3 . The catalytic absorptions
3 were better than the
non-catalytic absorptions. For MgCO3 , it was 10 g > 5 g > 7.5 g > 0 g for 1.0 mol/L, but 5 g > 0 g >
0.90
10 g > 7.5 g for 1.5 mol/L. We repeated the experiments of Figure 4A,B and Figure 6B at least twice.
Loading: mol CO2 / mol amine

0.80 and MgCO at large amount to 1.5 mol/L DEEA was probably due to the
This poorer effect of CaCO 3 3
“agglomeration” [40] of0.70
solid chemicals, where the liquid covered the solid surface area and inhibited
the catalysis. Such phenomena
0.60 were also reported by other researchers with 0 g > 50 g CaCO3 to
4.0 mol/L BEA + AMP amine blend [40]. For 1DMA-2P, it was 7.5 g > 5 g > 10 g > 0 g for 1.0 mol/L
0.50
10g CaCO3
0.40
7.5g CaCO3
0.30
5g CaCO3
0.20
0g CaCO3
Loading: mol CO2 / m

Loading: mol CO2 / mol


0.60 0.60
0.50 0.50
10g CaCO3 10g MgCO3
0.40 0.40
7.5g CaCO3 7.5g MgCO3
0.30 0.30
5g CaCO3 5g MgCO3
0.20 0.20
Molecules 2019, 24, 1009 0g CaCO3 0g MgCO3 11 of 20
0.10 0.10
0.00 0.00
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50
and 10 g > 7.5 g ≈ 5 g > 0 g for 1.5 mol/L for CaCO3 . For MgCO3Time , it was 10 g > 5 g > 7.5 g > 0 g
(min)
Time (min)
for 1.0 mol/L and 5 g > 7.5 g ≈ 10 g ≈ 0 g for 1.5 mol/L. The larger amount of MgCO3 at 1.5 mol/L
Figure
also resulted in 5. Catalytic CO2 absorption
agglomeration and made curves
theof catalysis
1.0 mol/L 1DMA-2P solvents atto293
comparable K and 1 atm. (A).absorption [40].
non-catalytic
CaCO3 0–10 g. (B). MgCO3 0–10 g.
The removal of agglomeration awaits further investigation.

Catalytic CO2 absorption of 1.5 mol/L 1DMA-2P


A
solvents with CaCO3
1.00
0.90
Loading: mol CO2 / mol amine

0.80
0.70
0.60
0.50
10g CaCO3
0.40
7.5g CaCO3
0.30
5g CaCO3
0.20
0g CaCO3
0.10
0.00
0 5 10 15 20 25 30 35 40 45 50
Molecules 2019, 24, x FOR PEER REVIEW 12 of 22
Time (min)

B Catalytic CO2 absorption of 1.5 mol/L 1DMA-2P


solvents with MgCO3
1.00
0.90
Loading: mol CO2 / mol amine

0.80
0.70
0.60
0.50
10g MgCO3
0.40
7.5g MgCO3
0.30
5g MgCO3
0.20
0g MgCO3
0.10
0.00
0 5 10 15 20 25 30 35 40 45 50
Time (min)

Catalytic
Figure 6. Figure CO2 CO
6. Catalytic absorption
2 absorptioncurves of1.51.5
curves of mol/L
mol/L 1DMA-2P
1DMA-2P solvents solvents at 1293
at 293 K and atm.K(A)
and 1 atm.
(A) CaCO3CaCO
0–103 0–10
g. (B) MgCO
g. (B) MgCO3 3 0–10
0–10 g.g.

The optimized amount of solid base chemicals is presented in Table 2. The orders were different
Under both amine concentrations without agglomeration, the catalytic absorptions were better
under different amine concentrations. For DEEA, it was 7.5 g > 10 g > 5 g > 0 g for 1.0 mol/L and 7.5 g
than the non-catalytic absorptions. For 1.0 mol/L, the absorption rates increased significantly with
≈ 5 g > 0 g > 10 g for 1.5 mol/L for CaCO3. The catalytic absorptions were better than the non-catalytic
different amounts
absorptions. For MgCOchemicals.
of solid 3, it was 10 g > For
5 g > 1.5
7.5 gmol/L, the
> 0 g for 1.0 catalytic
mol/L, but 5 g absorptions
> 0 g > 10 g > 7.5were better than
g for 1.5
mol/L. We repeated the experiments of Figure 4A,B and Figure 6B at least
non-catalytic cases at moderate ability. This difference was explained by Equations (13) and (14). twice. This poorer effect of
At a diluteCaCO 3 and MgCO3 at large amount to 1.5 mol/L DEEA was probably due to the “agglomeration” [40]
concentration of 1.0 mol/L the solid catalyst was helpful, for there are limited free R3 N
of solid chemicals, where the liquid covered the solid surface area and inhibited the catalysis. Such
amines around
phenomenaH2 CO 3 . also
were However,
reported atby a higher
other concentration
researchers with 0 g > 50 ofg 1.5
CaCO mol/L, there are more free R3 N
3 to 4.0 mol/L BEA + AMP
molecules amine
in solution with
blend [40]. Forhigher
1DMA-2P, pH itvalue
was 7.5 ingsolution,
> 5 g > 10 gthe
> 0 reaction rate and
g for 1.0 mol/L was10increased
g > 7.5 g ≈ 5with
g > 0[R3 N] and
g for 1.5 mol/L
the solid chemical hadforonlyCaCO 3. For MgCO
moderate 3, it was 10 g > 5 r
improvements gCO
> 7.5= g >k0RgNfor
[ R1.0
3 N mol/L
][ CO and
2 ] . 5 g > 7.5 g ≈ 10
2 3
g ≈ 0 g for 1.5 mol/L. The larger amount of MgCO3 at 1.5 mol/L also resulted in agglomeration and
made the catalysis comparable to non-catalytic absorption [40]. The removal of agglomeration awaits
further investigation.
Under both amine concentrations without agglomeration, the catalytic absorptions were better
than the non-catalytic absorptions. For 1.0 mol/L, the absorption rates increased significantly with
different amounts of solid chemicals. For 1.5 mol/L, the catalytic absorptions were better than non-
catalytic cases at moderate ability. This difference was explained by Equations (A1) and (A2). At a
Molecules 2019, 24, 1009 12 of 20

4.4. The Effect of Solid Base to CO2 Absorption to Tertiary Amine DEEA and 1DMA-2P with Comparison to
MEA and DEA
In addition to the periods of absorption profiles, the effect of CaCO3 and MgCO3 could also be
evaluated by the initial absorption rates, which was an important parameter [33]. The initial absorption
rates were shown in Figure 7 for non-catalytic absorption and optimized catalytic absorption. The effect
of solid alkalis was not the more the better, and there was an optimized mass. According to Section 2.2
with Equation (13), the increased mass of solid alkalis helped the proton transfer from H2 CO3 to solid
Molecules 2019, 24, x FOR PEER REVIEW 12 of 21
surface at start. However, after the optimized mass from Equation (14), the excess amount of solid
base inhibited
mol/L. We therepeated
protonthe transfer from of
experiments catalyst to R3and
Figure 4A,B N, Figure
and reduced the
6B at least overall
twice. This absorption
poorer effect rates.
of
At CaCO
optimized3 and MgCO 3 at largeboth
conditions, amount to 1.5
rates mol/L DEEA
increased was probably
significantly. Fordue to the “agglomeration”
DEEA, [40]
the initial absorption rate
of solid −chemicals,
2 where the liquid covered the solid surface
was 0.74 × 10 mol CO2 /min for 1.0 mol/L without catalysts, and it increased to 238% and 247% area and inhibited the catalysis. Such
phenomena were also reported by other researchers with 0 g > 50 g CaCO3 to 4.0 mol/L BEA + AMP
with the aid of CaCO3 and MgCO3 . The initial absorption rate increased to 1.39 × 10−2 mol CO2 /min
amine blend [40]. For 1DMA-2P, it was 7.5 g > 5 g > 10 g > 0g for 1.0 mol/L and 10g > 7.5 g ≈ 5 g > 0
for 1.5 mol/L,
g for 1.5but mol/Lincreased
for CaCOto only 122% and 135% with CaCO3 and
3. For MgCO3, it was 10 g > 5 g > 7.5 g > 0g
MgCO
for 1.0 mol/L3 .and
For51DMA-2P,
g > 7.5 g ≈ 10the initial
absorption 1.07 ×The − 2
g ≈ rate
0 g for was1.5 mol/L. 10 largermolamount
CO2 /min of MgCO for31.0 mol/L
at 1.5 mol/L without
also resultedcatalysts, and it increased
in agglomeration and to
153% and made 150% with CaCO
the catalysis comparable and MgCO
to .
non-catalyticThe initial
absorption absorption
[40]. The rate
removal increased
of to
agglomeration1.24 × 10
awaits −2 mol
3 3
CO2 /min further
for 1.5investigation.
mol/L and increased to 165% and 149% with CaCO3 and MgCO3 .
Under both amine concentrations without agglomeration, the catalytic absorptions were better
Compared with other studies of MEA and DEA, the absolute value of initial absorption rates
than the non-catalytic absorptions. For 1.0 mol/L, the absorption rates increased significantly with
of R3 N different
was smaller than MEA and DEA, because of the lower absorption rates and smaller second
amounts of solid chemicals. For 1.5 mol/L, the catalytic absorptions were better than non-
order rate constant
catalytic caseskat2 [27]. Withability.
moderate the aid ofdifference
This solid bases, wasthe initial by
explained rates properly
Equations increased.
13 and The effect of
14. At a dilute
solid chemicals
concentration was of stronger
1.0 mol/Latthe 1.0solid
mol/L andwas
catalyst turned moderate
helpful, at 1.5
for there are mol/L
limited free for
R3NDEEA. For 1DMA-2P,
amines around
H2COof
the increase 3. However, at a higherrates
initial absorption concentration
was similar of 1.5atmol/L, there are
the range more free R3for
of 150–165% N molecules
1.0 mol/L in solution
and 1.5 mol/L.
with higher pH value in solution, the reaction rate was increased
The overall absorption periods were reduced by about 46–48% at 1.0 mol/L for DEEA and 1DMA-2P, with [R 3N] and the solid chemical

had only moderate improvements r = k [R N][CO ].


but were reduced by only about 33% and 10% for 1.5 mol/L DEEA and 1DMA-2P.
Such
4.4.aThedifference was
Effect of Solid Basedue to 2the
to CO different
Absorption reaction
to Tertiary Aminemechanisms of different
DEEA and 1DMA-2P reactions.toFor CO2
with Comparison
reactionMEAwithand tertiary
DEA amine, it was the based catalyzed hydration mechanism. The R3 N do not react
with CO2 directly,
In addition butto accept
the periods protons from [H
of absorption 2 CO3 ].theThe
profiles, stronger
effect of CaCO3basicity
and MgCO of3 the solvents
could also be led to
strongerevaluated
proton affinity, and caused
by the initial absorptionbetter CO2which
rates, absorption.
was an On the basis
important of kinetic
parameter studies,
[33]. the second
The initial
order rate constantrates
absorption (k2 ) were
was related
shown in to pKa
Figure of7thefortertiary amine.absorption
non-catalytic [5] The increased aminecatalytic
and optimized concentration
led to a absorption.
bigger pHThe valueeffect
ofofthesolid alkalis was
solution andnot the more
higher the better,rates.
absorption and there
Thewas an optimized
solid mass. could
base chemicals
According to Section − 2.2 with Equation 13, the increased mass
not directly affect the [OH ] or pH value in solvent, and provided only moderate enhancement of CO2of solid alkalis helped the proton
transfer from H2CO3 to solid surface at start. However, after the optimized mass from Equation 14,
absorption rates at higher concentrations. In short, the solid alkalis were effective for tertiary amines,
the excess amount of solid base inhibited the proton transfer from catalyst to R3N, and reduced the
but the overall
effectsabsorption
were not rates.
as good as for primary and secondary amines. They were more effective at a
dilute concentration.
At optimized conditions, both rates increased significantly. For DEEA, the initial absorption rate
However,
was 0.74 ×for 10−2MEA
mol CO and2/min DEA,
for 1.0the COwithout
mol/L 2 reaction is driven
catalysts, and it by Zwitterion
increased to 238%mechanism
and 247% with with the
the aid of CaCO 3 and MgCO3. The initial absorption rate increased to 1.39 × 10−2 mol CO2/min for 1.5
carbamate formation as products [28]. Solid alkalis might enhance the mass transfer or reduce
mol/L, but increased to only 122% and 135% with CaCO3 and MgCO3. For 1DMA-2P, the initial
the activation energy (Ea) of the−2 reaction process and facilitate N-C bond formation of CO2 -amine.
absorption rate was 1.07 × 10 mol CO2/min for 1.0 mol/L without catalysts, and it increased to 153%
The solid surface area contains abundant active sites which facilitate CO absorption in another reaction
and 150% with CaCO3 and MgCO3. The initial absorption rate increased2to 1.24 × 10−2 mol CO2/min
pathway for[19,20].
1.5 mol/L and increased to 165% and 149% with CaCO3 and MgCO3.

a 3.0 1.0 mol/L DEEA

2.22
(×10-2 mol CO2 / min)

2.0 1.76 1.66 1.63


1.31
Iabs_rate

1.0 0.74 0.80 0.74

0.0
CaCO3 MgCO3
0g 5g 7.5 g 10 g

Figure 7. Cont.
Molecules 2019, 24, 1009 13 of 20
Molecules 2019, 24, x FOR PEER REVIEW 13 of 21

1.5 mol/L DEEA


b 2.5

1.88
(×10-2 mol CO2 / min)
2.0 1.70
1.59 1.58
1.39 1.39
1.5 1.28
Iabs_rate

1.09
1.0

0.5

0.0
CaCO3 MgCO3
0g 5g 7.5g 10g

1.0 mol/L 1DMA-2P


c 2.0
1.64 1.61 1.56
mol CO2 / min)

1.5
1.07 1.16 1.13 1.07 1.15
Iabs_rate

1.0

0.5
(×10-2

0.0
CaCO3 MgCO3
0g 5g 7.5g 10g

1.5 mol/L 1DMA-2P


d 2.5
1.87
mol CO2 / min)

2.0 1.88 1.69 1.75


1.47 1.42
1.5 1.24 1.24
Iabs_rate

1.0
(×10-2

0.5

0.0
CaCO3 MgCO3
0g 5g 7.5g 10g

Figure 7. Initial
Initial Absorption
Absorption rates
rates forfor DEEAand
DEEA and 1DMA-2P
1DMA-2P with optimized
with amount
optimized of CaCO
amount and
of 3CaCO
Figure 7. 3 and
MgCO3, from 1.0–1.5 mol/L at 293K and 1 atm. (a) 1.0 mol/L DEEA, (b) 1.5 mol/L DEEA, (c) 1.0 mol/L
MgCO3 , from 1.0–1.5 mol/L at 293 K and 1 atm. (a) 1.0 mol/L DEEA, (b) 1.5 mol/L DEEA, (c) 1.0 mol/L
1DMA-2P, (d) 1.5 mol/L 1DMA-2P.
1DMA-2P, (d) 1.5 mol/L 1DMA-2P.
Compared with other studies of MEA and DEA, the absolute value of initial absorption rates of
4.5. The RCO 2 Desorption Profiles with Heat Duty Analyses
3N was smaller than MEA and DEA, because of the lower absorption rates and smaller second order

Therate
CO constant k2 [27]. With the aid of solid bases, the initial rates properly increased. The effect of solid
2 desorption profiles were plotted in Figures 8 and 9 for 1~2 mol/L DEEA and 1DMA-2P
chemicals was stronger at 1.0 mol/L and turned moderate at 1.5 mol/L for DEEA. For 1DMA-2P, the
solvents. This range of amine concentration was suitable for industrial application, for the 0–2 mol/L
increase of initial absorption rates was similar at the range of 150–165% for 1.0 mol/L and 1.5 mol/L.
MDEA The were usually
overall blended
absorption with
periods 5Mreduced
were MEA to prepare
by about MEA-R
46–48% N solvents,
at 1.03mol/L for DEEAusually 5 + 0.5, 5 + 1,
and 1DMA-2P,
5 + 1.25,but
5 +were
1.5 M [6,7,9].
reduced byThe
onlyoperation
about 33% condition
and 10% forof1.5
MEA-R 3 N was
mol/L DEEA also
and 0.20~0.60 mol/mol. Moreover,
1DMA-2P.
for 1DMA-2P, the solubility was low. Small amounts of crystal were observed in 2.0 mol/L solvent at
293 K, but it was soluble in water at 363 K.
The CO2 desorption profiles were plotted in Figures 8 and 9 for 1~2 mol/L DEEA and 1DMA-2P
solvents. This range of amine concentration was suitable for industrial application, for the 0–2 mol/L
MDEA were usually blended with 5M MEA to prepare MEA-R3N solvents, usually 5 + 0.5, 5 + 1, 5 +
1.25, 5 + 1.5 M [6,7,9]. The operation condition of MEA-R3N was also 0.20~0.60 mol/mol. Moreover,
Molecules 2019,
for 24, 1009
1DMA-2P, 14 ofat
the solubility was low. Small amounts of crystal were observed in 2.0 mol/L solvent 20
293 K, but it was soluble in water at 363 K.

The CO2 desorption curves of 1.0 mol/L The CO2 desorption curves of 1.0 mol/L
a DEEA solvents with γ-Al2O3 b DEEA solvents with HZSM-5
1.00 1.00
blank blank

CO2 loading (mol CO2 / mol


CO2 loading (mol CO2 / mol

0.80 5.0g γ-Al2O3 0.80 5.0g H-ZSM-5


7.5g γ-Al2O3 7.5g H-ZSM-5
0.60 10.0g γ-Al2O3 0.60 10.0g H-ZSM-5

amine)
amine)

0.40 0.40

0.20 0.20

0.00 0.00
0 60 120 180 240 300 360 420 480 0 60 120 180 240 300 360 420 480
Time (min) Time (min)

The CO2 desorption curves of 1.5 mol/L d The CO2 desorption curves of 1.5
c
DEEA solvents with γ-Al2O3 mol/L DEEA solvents with H-ZSM-5
1.00 1.00
CO2 loading (mol CO2 / mol

CO2 loading (mol CO2 / mol


blank blank
5.0g γ-Al2O3 0.80 5.0g H-ZSM-5
0.80
7.5g γ-Al2O3 7.5g H-ZSM-5
amine)

amine)

0.60 10.0g γ-Al2O3 0.60 10.0g H-ZSM-5

0.40 0.40

0.20 0.20

0.00 0.00
0 60 120x FOR
180PEER
240 300 360 420 0 60 120 180 240 300 360 420
Molecules 2019, 24, REVIEW 16 of 22
Time (min) Time (min)

The CO2 desorption curves of 2.0 mol/L The CO2 desorption curves of 2.0 mol/L
e f
DEEA solvents with γ-Al2O3 DEEA solvents with HZSM-5
1.00 1.00
blank blank
CO2 loading (mol CO2 / mol

CO2 loading (mol CO2 / mol

5.0g γ-Al2O3 5.0g H-ZSM-5


0.80 0.80
7.5g γ-Al2O3 7.5g H-ZSM-5
0.60 10.0g γ-Al2O3 10.0g H-ZSM-5
amine)

0.60
amine)

0.40 0.40

0.20 0.20

0.00 0.00
0 60 120 180 240 300 360 0 60 120 180 240 300 360
Time (min) Time (min)

Figure 8. Catalytic CO2 desorption


Figure 8. Catalytic curves
CO2 desorption of 1.0–2.0
curves mol/L
of 1.0–2.0 mol/LDEEA
DEEAsolvents
solvents at 363
363KKand
and1 1atm,
atm, with
with 0–
10 g γ-Al O and H-ZSM-5. (a,b) 1.0 mol/L with γ-Al O and HZSM-5; (c,d) 1.5
0–10 g γ-Al2 O3 and H-ZSM-5. (a,b) 1.0 mol/L with γ-Al2 O3 and HZSM-5; (c,d) 1.5 mol/L with γ-Al22O
2 3 2 3 mol/L with γ-Al O33
and HZSM-5;
and HZSM-5; (e,f) 2.0with
(e,f) 2.0 mol/L mol/Lγ-Al
with2 O
γ-Al 2O3 and HZSM-5.
3 and HZSM-5.

The CO2 desorption curves of 1.0 mol/L The CO2 desorption curves of 1.0 mol/L
a b 1DMA-2P solvents with HZSM-5
1DMA-2P solvents with γ-Al2O3 1.00
1.00
CO2 loading (mol CO2 / mol

blank blank
CO2 loading (mol CO2 / mol

0.80 0.80
5.0g γ-Al2O3 5.0g H-ZSM-5
amine)

0.60 7.5g γ-Al2O3 0.60 7.5g H-ZSM-5


amine)

10.0g γ-Al2O3 10.0g H-ZSM-5


0.40 0.40

0.20 0.20

0.00 0.00
0 30 60 90 120 150 180 210 240 0 30 60 90 120 150 180 210 240
Time (min) Time (min)
C
0.00 0.00
0 60 120 180 240 300 360 0 60 120 180 240 300 360
Time (min) Time (min)

Figure 8. Catalytic CO2 desorption curves of 1.0–2.0 mol/L DEEA solvents at 363 K and 1 atm, with 0–
Molecules 2019, 24, 1009
10 g γ-Al2O3 and H-ZSM-5. (a,b) 1.0 mol/L with γ-Al2O3 and HZSM-5; (c,d) 1.5 mol/L with γ-Al2O3 15 of 20
and HZSM-5; (e,f) 2.0 mol/L with γ-Al2O3 and HZSM-5.

The CO2 desorption curves of 1.0 mol/L The CO2 desorption curves of 1.0 mol/L
a b 1DMA-2P solvents with HZSM-5
1DMA-2P solvents with γ-Al2O3 1.00
1.00
CO2 loading (mol CO2 / mol

blank blank

CO2 loading (mol CO2 / mol


0.80 0.80
5.0g γ-Al2O3 5.0g H-ZSM-5
amine)

0.60 7.5g γ-Al2O3 0.60 7.5g H-ZSM-5

amine)
10.0g γ-Al2O3 10.0g H-ZSM-5
0.40 0.40

0.20 0.20

0.00 0.00
0 30 60 90 120 150 180 210 240 0 30 60 90 120 150 180 210 240
Time (min) Time (min)

c The CO2 desorption curves of 1.5 mol/L d The CO2 desorption curves of 1.5 mol/L
1DMA-2P solvents with γ-Al2O3 1DMA-2P solvents with H-ZSM-5
1.00 1.00

CO2 loading (mol CO2 / mol


blank blank
CO2 loading (mol CO2 / mol

0.80 5.0g γ-Al2O3 0.80 5.0g H-ZSM-5

7.5g γ-Al2O3 7.5g H-ZSM-5


amine)
0.60 0.60
amine)

10.0g γ-Al2O3 10.0g H-ZSM-5


0.40 0.40

0.20 0.20

0.00 0.00
0
Molecules 30 24,60
2019, 90 REVIEW
x FOR PEER 120 150 180 0 30 60 90 120 150 17 of180
22
Time (min) Time (min)

e The CO2 desorption curves of 2.0 mol/L The CO2 desorption curves of 2.0 mol/L
1DMA-2P solvents with γ-Al2O3 f
1.00 1DMA-2P solvents with HZSM-5
1.00
CO2 loading (mol CO2 / mol

blank blank
CO2 loading (mol CO2 / mol

0.80 5.0g γ-Al2O3 0.80 5.0g H-ZSM-5

0.60 7.5g γ-Al2O3 7.5g H-ZSM-5


amine)

0.60
amine)

10.0g γ-Al2O3 10.0g H-ZSM-5


0.40 0.40

0.20 0.20

0.00 0.00
0 60 120 180 0 60 120 180
Time (min) Time (min)

Figure 9. Catalytic
Figure 9.CO 2 desorption
Catalytic curvescurves
CO2 desorption of 1.0–2.0 mol/L
of 1.0–2.0 mol/L1DMA-2P solventsat at
1DMA-2P solvents 363363 K and
K and 1 atm,1 atm,
with γ-Al2 O3with
andγ-Al
H-ZSM-5. (a,b) 1.0(a,b)
2O3 and H-ZSM-5. mol/L with γ-Al
1.0 mol/L O
with γ-Al
2 3 2 and
O 3 HZSM-5;
and HZSM-5;(c,d)
(c,d) 1.5
1.5 mol/L
mol/L with
with γ-Al
γ-Al 2 O 3 2 O3

and HZSM-5; and HZSM-5;


(e,f) (e,f) 2.0
2.0 mol/L mol/L
with withO
γ-Al γ-Al 2O3 and HZSM-5.
and HZSM-5.
2 3

From the CO2 desorption curves, there were some clues. For DEEA, the catalytic desorption was
From the CO 2 desorption curves, there were some clues. For DEEA, the catalytic desorption was
better than the non-catalytic one, with the order of H-ZSM-5 > γ-Al2O3 > non-catalyst at 5.0 g, 7.5 g
better than the non-catalytic one, with the order of H-ZSM-5 > γ-Al2 O≈3 γ-Al
and 10 g through 1.0~2.0 mol/L. For 1DMA-2P, the effects of H-ZSM-5
> non-catalyst at 5.0 g, 7.5 g
2O3 > non-catalyst at 5.0
and 10 g through
and 7.51.0~2.0 mol/L.>For
g, but H-ZSM-5 γ-Al1DMA-2P, the effects
2O3 > non-catalyst at 10 gof
at H-ZSM-5 ≈ γ-Alof
optimized amount 2O 3 > non-catalyst
H-ZSM-5. This effectat 5.0
and 7.5 g, but
wasH-ZSM-5
reasonable>because
γ-Al2 Othe
3 >CO
non-catalyst at 10 gfrom
2 loading decreased at optimized amountatof
0.80 to 0.30 mol/mol theH-ZSM-5.
first 30 min.This
Thereeffect
were abundant bicarbonates [HCO −], and more [R3NH+] from ion speciation plot, which made the
was reasonable because the CO2 loading decreased from 0.80 to 0.30 mol/mol at the first 30 min. There
3

CO2 production comparable −


were abundant bicarbonates [HCO3 with γ-Al2O3 and HZSM-5.
], and more [R3 NH+ ] from Both solid acids could enhance the proton
ion speciation plot, which made the
transfer process. However, at 10 g H-ZSM-5 of the series, an excess amount of H-ZSM-5 provided
CO2 production comparable with γ-Al
excessive protons into the solvents,
O and HZSM-5. Both solid
2 3and then fully reacted with [HCO acids could enhance the proton
3−] and produced more CO2 to
transfer process.
reduce However,
heat duty. at 10 g H-ZSM-5 of the series, an excess amount of H-ZSM-5 provided
Time (min) Time (min)

Figure 9. Catalytic CO2 desorption curves of 1.0–2.0 mol/L 1DMA-2P solvents at 363 K and 1 atm,
with γ-Al2O3 and H-ZSM-5. (a,b) 1.0 mol/L with γ-Al2O3 and HZSM-5; (c,d) 1.5 mol/L with γ-Al2O3
and HZSM-5; (e,f) 2.0 mol/L with γ-Al2O3 and HZSM-5.
Molecules 2019, 24, 1009 16 of 20
From the CO2 desorption curves, there were some clues. For DEEA, the catalytic desorption was
better than the non-catalytic one, with the order of H-ZSM-5 > γ-Al2O3 > non-catalyst at 5.0 g, 7.5 g
and 10
excessive g through
protons into1.0~2.0 mol/L. For
the solvents, and1DMA-2P, thereacted
then fully effects ofwith
H-ZSM-5 − ] and
[HCO≈3γ-Al 2O3 > non-catalyst
produced moreat 5.0
CO2 to
and 7.5 g,
reduce heat duty. but H-ZSM-5 > γ-Al 2O3 > non-catalyst at 10 g at optimized amount of H-ZSM-5. This effect

was reasonable because the CO2 loading decreased from 0.80 to 0.30 mol/mol at the first 30 min. There
The heat duty for the first 30 min−was calculated in +Figures 10 and 11. The heat duty was mostly
were abundant bicarbonates [HCO3 ], and more [R3NH ] from ion speciation plot, which made the
determined by the CO2 production (nCO2 ) as the heat inputs were similar for the first 30 min. During
CO2 production comparable with γ-Al2O3 and HZSM-5. Both solid acids could enhance the proton
that period,
transfermost of the
process. CO2 desorption
However, processof
at 10 g H-ZSM-5 was
thecompleted as theamount
series, an excess CO2 loading decreased
of H-ZSM-5 from 0.80
provided
to 0.30 mol/mol.
excessive The into
protons CO2thedesorption curves
solvents, and thendid
fullynot shift with
reacted significantly after
[HCO3−] and loadingmore
produced <0.30COmol/mol,
2 to

sincereduce
most [HCO − ] was exhausted.
heat duty.
3

Molecules 2019, 24, x FOR PEER REVIEW 18 of 22

that period, most of the CO2 desorption process was completed as the CO2 loading decreased from
0.80 to 0.30 mol/mol. The CO2 desorption curves did not shift significantly after loading <0.30
Figure
Figure
mol/mol, 10.
10. The
since The
most heat
heat duty
duty
[HCO of of
thethe DEEAatatfirst
DEEA
3−] was exhausted.
first 30
30 min
min from
from1.0
1.0toto2.0
2.0mol/L
mol/Lat 363 K and
at 363 1 atm.
K and 1 atm.

The heat duty for the first 30 min was calculated in Figures 10 and 11. The heat duty was mostly
determined by the CO2 production (nCO2) as the heat inputs were similar for the first 30 min. During

Figure 11. The heat duty of 1DMA-2P at first 30 min from 1.0 to 2.0 mol/L at 363 K and 1 atm.
Figure 11. The heat duty of 1DMA-2P at first 30 min from 1.0 to 2.0 mol/L at 363 K and 1 atm.

Based on Table 3, it was discovered that the heat duty was properly reduced. For γ-Al2 O3 ,
Based on Table 3, it was discovered that the heat duty was properly reduced. For γ-Al2O3, the
the order
orderwas
was7.5 10 ≈
7.5gg>>10 ≈ 55 gg>>00g,g,and
andfor
forH-ZSM-5,
H-ZSM-5, thethe order
order waswas 10>g7.5
10 g > 7.5
g>g5g> >5 0gg.
> At
0 g.optimized
At optimized
catalytic conditions,
catalytic the heat
conditions, dutyduty
the heat was was
reduced by about
reduced 83-98
by about % under
83-98 % underdifferent conditions.
different conditions.ForFor
DEEA,
the reduction
DEEA, the ofreduction
heat dutyoffollowed
heat dutythe order of
followed the1.5 > 1.0of>1.5
order 2.0>mol/L. For
1.0 > 2.0 1DMA-2P,
mol/L. the reduction
For 1DMA-2P, the of
heat duty followed
reduction of heatthe order
duty of 1.5the
followed > 2.0 > 1.0
order mol/L.
of 1.5 > 2.0 >Hence, theHence,
1.0 mol/L. aminetheconcentration was preferred
amine concentration was at
preferred at 1.0–1.5 mol/L for DEEA and at 1.5–2.0 mol/L
1.0–1.5 mol/L for DEEA and at 1.5–2.0 mol/L for 1DMA-2P with catalysts. for 1DMA-2P with catalysts.

Table 3. The relative heat duty (%) of different amine solvents under optimized catalysis at 363 K and
1 atm.

Heat Duty (kJ/mol CO2)


Amine Solvents Optimized Catalysts
Optimized Catalysis Non-Catalyst Ratio (%)
DEEA 7.5 g γ-Al2O3 1741.8 2004.2 86.91%
Molecules 2019, 24, 1009 17 of 20

Table 3. The relative heat duty (%) of different amine solvents under optimized catalysis at 363 K and 1 atm.

Heat Duty (kJ/mol CO2 )


Amine Solvents Optimized Catalysts
Optimized Catalysis Non-Catalyst Ratio (%)
DEEA 7.5 g γ-Al2 O3 1741.8 2004.2 86.91%
1.0 mol/L 10 g HZSM-5 1685.7 2004.2 84.11%
7.5 g γ-Al2 O3 1032.8 1136.9 90.84%
1.5 mol/L
10 g HZSM-5 945.0 1136.9 83.12%
7.5 g γ-Al2 O3 640.3 703.4 91.03%
2.0 mol/L
10 g HZSM-5 606.5 703.4 86.22%
1DMA-2P 10 g γ-Al2 O3 1401.3 1427.0 98.20%
1.0 mol/L 10 g HZSM-5 1352.6 1427.0 94.78%
7.5 g γ-Al2 O3 816.3 1040.2 78.48%
1.5 mol/L
10 g HZSM-5 760.5 1040.2 73.11%
10 g γ-Al2 O3 545.0 616.5 88.40%
2.0 mol/L
10 g HZSM-5 522.1 616.5 84.69%

4.6. The Effect of Solid Acid to Tertiary Amine DEEA and 1DMA-2P and Compared with MEA and DEA
The effect of solid acids to 1DMA-2P and DEEA was shown in Table 3. Briefly, the effects
of H-ZSM-5 and γ-Al2 O3 were different due to different reaction schemes and mechanisms. For
tertiary amine, HZSM-5 was better than γ-Al2 O3 for DEEA within the scope of 1–2 mol/L and 0–10
g. For 1DMA-2P, H-ZSM-5 was comparable to γ-Al2 O3 with inadequate catalysis (5 g and 7.5 g),
and better than γ-Al2 O3 at 10 g. The increased mass of H-ZSM-5 led to better desorption performance,
with the order of: 10 g > 7.5 g > 5 g > 0 g.
This could be explained by the reactions (18) and (19) in Section 2.3. The Equation (18) reflected
the fact that CO2 desorption rates were determined by mass of H-ZSM-5. Since H-ZSM-5 was a proton
donor that reacted with bicarbonate right away, more H-ZSM-5 led to better desorption performance.
After loading < 0.25 mol/mol, the effect of H-ZSM-5 was almost the same despite different masses
because [HCO3 − ] was exhausted from ion speciation plot [17].
On the other hand, the effect of γ-Al2 O3 was complicated. In some cases, 10 g was the best, and in
other cases, 7.5 g was the best, which was even better than 10 g and 5 g. This trend was quite different
from MEA [21] and DEA [26]. In short, for single catalysts, the increased mass of solid acid led to
better desorption performance for MEA [21] and DEA [26]. From Table S1 [26] it’s concluded that
the order was H-ZSM-5 > blended catalyst (γ-Al2 O3 /H-ZSM-5) > γ-Al2 O3 at 0.50–0.30 mol/mol for
MEA, blended catalyst (γ-Al2 O3 /H-ZSM-5) > γ-Al2 O3 > H-ZSM-5 at 0.30–0.15 mol/mol for MEA [21].
For DEA, HZSM-5 was better than γ-Al2 O3 in both rich and lean regions [26]. However, excessive
amounts of γ-Al2 O3 might reduce the catalysis of tertiary amines from optimum dose, while the effect
was still better than that of non-catalyst.
The reason was based on reactions (15)–(17), for the role of γ-Al2 O3 was twofold. It had to accept
[H+ ] from [R3 NH+ ] first because it contained no proton, and then released [H+ ] to [HCO3 − ]. The release
of CO2 was determined by Equation (17), and it was affected by [HAlO2 ]. From Equation (16), HAlO2
was generated from proton release from [R3 NH+ ] to [AlO2 − ]. From Equation (15), the concentration
of [AlO2 − ] increased with increased amounts of γ-Al2 O3 . The increased [AlO2 − ] enhanced the
acceptance of [H+ ] in Equation (16), but the extra [AlO2 − ] might inhibit the proton transfer to [HCO3 − ]
in Equation (17) and then affected the CO2 desorption rates. The increased mass of γ-Al2 O3 enhanced
desorption firstly, and then reduce desorption after the optimum, for the excessive amounts of [AlO2 − ]
might inhibit the proton transfer from [HAlO2 ] to [HCO3 − ]. Similar to solid base (CaCO3 and MgCO3 )
to CO2 absorption, there was also an optimized dose of γ-Al2 O3 , and inadequate or excessive amounts
of it would reduce the effects of CO2 desorption.
Molecules 2019, 24, 1009 18 of 20

5. Conclusions
(1). The CO2 equilibrium solubility and pKa of DEEA and 1DMA-2P were comparable to
published data, and the scope was expanded to 293 K from the previous 298–313 K.
(2). The existence of CaCO3 and MgCO3 as solid alkalis accelerated the CO2 -R3 N absorption, via
the “base-catalyzed mechanism”. The effect of solid alkaline was indirect, it facilitated proton transfer
from H2 CO3 firstly, and the proton was then transferred to stronger base R3 N. The increased mass of
solid base boosted proton transfer, but the excess amount might inhibit the proton transfer from solid
to R3 N. Therefore, there was an optimized dose of CaCO3 and MgCO3 as catalysts. For solid alkalis,
their effects were significant at 1.0 mol/L, but moderate at 1.5 mol/L because the increase of the amine
concentration resulted in the increase of absorption rates and the increase of pH value. High amine
concentration provided more free molecules into the solution to enhance the proton transfer of H2 CO3 .
(3). The solid acids could enhance the CO2 desorption and reduce the heat duties of both tertiary
amines. The effect of catalytic desorption was better than that of non-catalytic ones. For DEEA, it was
H-ZSM-5 > γ-Al2 O3 > non-catalyst. For 1DMA-2P, it was H-ZSM-5 ≈ γ-Al2 O3 > non-catalyst with
inadequate catalysts, but H-ZSM-5 > γ-Al2 O3 > non-catalyst at optimized performance.
(4). The effect of Bronsted acid/proton donor H-ZSM-5 to CO2 desorption was straightforward,
that is, the more the better as it reacts with bicarbonate directly. The effect of Lewis acid such as
γ-Al2 O3 to CO2 desorption was indirect. The increased mass of γ-Al2 O3 resulted in increased [AlO2 − ],
which could boost proton transfer of [R3 NH+ ] to generate [HAlO2 ]. However, the excess amount of
[AlO2 − ] might inhibit the proton transfer of [HAlO2 ] to [HCO3 − ] and release CO2 . Therefore, there
was surely an optimized dose of γ-Al2 O3 , which was not the more the better.

Supplementary Materials: The following are available online. Figure S1: Stirred Cell Reactor for CO2 -Amine
interactions with a water scrubbing process. Figure S2: The schematic diagram of the CO2 desorption process
with oil bath. Figure S3: The CO2 equilibrium solubility of investigated amines and our graphical method of
solubility test. Figure S4: The SEM and BET of pelletized CaCO3 and MgCO3 . Table S1. The order of catalysis for
MEA, DEA and MEA+DEA blended amines under different cases of blended solvents.
Author Contributions: H.S. completed the whole process of the literature survey, manuscript writing,
and experiment designs. S.Z. analyzed the chemical reactions from the mechanism of CO2 -R3 N on a molecular
level. L.C. helped to complete the catalysis part. M.H. completed the CO2 absorption experiments of DEEA and
1DMA-2P, and Q.W. completed the calculation of CO2 initial absorption rates. L.Z. completed the desorption
process, and L.C. completed the calculation of heat duty study. P.T. helped the revision and proofreading.
Funding: The National Natural Science Foundation of China (NSFC Nos. 21606150) are gratefully acknowledged.
It is also supported by Young Eastern Scholar with complement (QD 2016011, QD 2016011-005-011). Natural
Sciences and Engineering Research Council of Canada (NSERC) are also acknowledged.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Afkhamipour, M.; Mofarahi, M. A modeling-optimization framework for assessment of CO2 absorption
capacity by novel amine solutions: 1DMA2P, 1DEA2P, DEEA, and DEAB. J. Clean. Prod. 2018, 171, 234–249.
2. Liang, Z.; Rongwong, W.; Liu, H.; Fu, K.; Gao, H.; Cao, F.; Zhang, R.; Sema, T.; Henni, A.; Sumon, K.; et al.
Recent progress and new developments in post-combustion carbon-capture technology with amine based
solvents. Int. J. Greenh. Gas Control 2015, 40, 26–54. [CrossRef]
3. Liu, H.; Xiao, M.; Tontiwachwuthikul, P.; Liang, Z. A Novel Model for Correlation and Predication of the
Equilibrium CO2 Solubility in Seven Tertiary Solvents. Energy Procedia 2017, 105, 4476–4481. [CrossRef]
4. Mofarahi, M.; Khojasteh, Y.; Khaledi, H.; Farahnak, A. Design of CO2 absorption plant for recovery of CO2
from flue gases of gas turbine. Energy 2008, 33, 1311–1319. [CrossRef]
5. Min Xiao, H.L. Raphael Idem, Paitoon Tontiwachwuthikul, Zhiwu Liang, A study of structure–activity
relationships of commercial tertiary amines for post-combustion CO2 capture. Appl. Energy 2016, 184,
219–229. [CrossRef]
Molecules 2019, 24, 1009 19 of 20

6. Shi, H.C.; Naami, A.; Idem, R.; Tontiwachwuthikul, P. Catalytic and non catalytic solvent regeneration
during absorption-based CO2 capture with single and blended reactive amine solvents. Int. J. Greenh. Gas
Control 2014, 26, 39–50. [CrossRef]
7. Srisang, W.; Pouryousefi, F.; Osei, P.A.; Decardi-Nelson, B.; Akachuku, A.; Tontiwachwuthikul, P.; Idem, R.
CO2 capture efficiency and heat duty of solid acid catalyst-aided CO2 desorption using blends of
primary-tertiary amines. Int. J. Greenh. Gas Control 2018, 69, 52–59. [CrossRef]
8. Srisang, W.; Pouryousefi, F.; Osei, P.A.; Decardi-Nelson, B.; Akachuku, A.; Tontiwachwuthikul, P.; Idem, R.
Evaluation of the heat duty of catalyst aided amine post combustion CO2 capture. Chem. Eng. Sci. 2017, 170,
48–57. [CrossRef]
9. Liu, H.L.; Zhang, X.; Gao, H.X.; Liang, Z.W.; Idem, R.; Tontiwachwuthikul, P. Investigation of CO2
Regeneration in Single and Blended Amine Solvents with and without Catalyst. Ind. Eng. Chem. Res.
2017, 56, 7656–7664. [CrossRef]
10. Luo, X.; Chen, N.; Liu, S.; Rongwong, W.; Idem, R.O.; Tontiwachwuthikul, P.; Liang, Z. Experiments and
modeling of vapor-liquid equilibrium data in DEEA-CO2 -H2 O system. Int. J. Greenh. Gas Control 2016, 53,
160–168. [CrossRef]
11. Rayer, A.V.; Sumon, K.Z.; Jaffari, L.; Henni, A. Dissociation Constants (pKa) of Tertiary and Cyclic Amines:
Structural and Temperature Dependences. J. Chem. Eng. Data 2014, 59, 3805–3813. [CrossRef]
12. Xu, B.; Gao, H.; Luo, X.; Liao, H.; Liang, Z. Mass transfer performance of CO2 absorption into aqueous DEEA
in packed columns. Int. J. Greenh. Gas Control 2016, 51, 11–17. [CrossRef]
13. Kadiwala, S.; Rayer, A.V.; Henni, A. Kinetics of carbon dioxide (CO2 ) with ethylenediamine,
3-amino-1-propanol in methanol and ethanol, and with 1-dimethylamino-2-propanol and 3-dimethylamino-1-
propanol in water using stopped-flow technique. Chem. Eng. J. 2012, 179, 262–271. [CrossRef]
14. Liang, Y.J.; Liu, H.L.; Rongwong, W.; Liang, Z.W.; Idem, R.; Tontiwachwuthikul, P. Solubility, absorption
heat and mass transfer studies of CO2 absorption into aqueous solution of 1-dimethylamino-2-propanol.
Fuel 2015, 144, 121–129. [CrossRef]
15. Liu, H.; Gao, H.; Idem, R.; Tontiwachwuthikul, P.; Liang, Z. Analysis of CO2 solubility and absorption heat
into 1-dimethylamino-2-propanol solution. Chem. Eng. Sci. 2017, 170, 3–15. [CrossRef]
16. Hamborg, E.S.; Versteeg, G.F. Dissociation Constants and Thermodynamic Properties of Amines and
Alkanolamines from (293−353) K. Chem. Eng. Data 2009, 54, 1318–1328. [CrossRef]
17. Liu, H.; Idem, R.; Tontiwachwuthikul, P.; Liang, Z. Study of Ion Speciation of CO2 Absorption into Aqueous
1-Dimethylamino-2-propanol Solution Using the NMR Technique. Ind. Eng. Chem. Res. 2017, 56, 8697–8704.
[CrossRef]
18. Narku-Tetteh, J.; Afari, D.B.; Coker, J.; Idem, R. Evaluation of the Roles of Absorber and Desorber Catalysts
in the Heat Duty and Heat of CO2 Desorption from Butylethanolamine–2-Amino-2-methyl-1-propanol
and Monoethanolamine–Methyldiethanolamine Solvent Blends in a Bench-Scale CO2 Capture Pilot Plant.
Energy Fuels 2018, 32, 9711–9726. [CrossRef]
19. Idem, R.; Shi, H.; Gelowitz, D.; Tontiwachwuthikul, P. Catalytic Method and Apparatus for Separation
Gaseous Component from an Incoming Gas Stream. U.S. Patent US 2013/0108532 A1, 2 May 2013.
20. Shi, H.C.; Zhou, Y.L.; Zuo, Y.H.; Cui, L.F.; Idem, R.; Tontiwachwuthikul, P. Heterogeneous
catalysis of CO2 -diethanolamine absorption with MgCO3 and CaCO3 and comparing to non-catalytic
CO2 -monoethanolamine interactions. React. Kinet. Mech. Catal. 2017, 122, 539–555. [CrossRef]
21. Liang, Z.W.; Idem, R.; Tontiwachwuthikul, P.; Yu, F.H.; Liu, H.L.; Rongwong, W. Experimental study on the
solvent regeneration of a CO2 -loaded MEA solution using single and hybrid solid acid catalysts. AIChE J.
2016, 62, 753–765. [CrossRef]
22. Shi, H.C.; Zhou, Y.L.; Si, M.Y.; Zuo, Y.H.; Kang, S.F.; Wang, Y.G.; Cui, L.F.; Idem, R.; Tontiwachwuthikul, P.;
Liang, Z.W. Amine Regeneration tests on MEA, DEA and MMEA with respect to carbamate stability analyses.
Can. J. Chem. Eng. 2017, 95, 1471–1479. [CrossRef]
23. Decardi-Nelson, B.; Akachuku, A.; Osei, P.; Srisang, W.; Pouryousefi, F.; Idem, R. A flexible and robust
model for low temperature catalytic desorption of CO2 from CO2 -loaded amines over solid acid catalysts.
Chem. Eng. Sci. 2017, 170, 518–529. [CrossRef]
24. Bhatti, U.H.; Shah, A.K.; Kim, J.N.; You, J.K.; Choi, S.H.; Lim, D.H.; Nam, S.; Park, Y.H.; Baek, H. Effects
of transition metal oxide catalysts on MEA solvent regeneration for the post-combustion carbon capture
process. ACS Sustain. Chem. Eng. 2017, 5, 5862–5868. [CrossRef]
Molecules 2019, 24, 1009 20 of 20

25. Qinghua, L.; Sam, T.; Mohammed, A.A.; Huaigang, C.; Armistead, G.R.; Hertanto, A.; Maciej, R.; Maohong, F.
Catalyst-TiO(OH)2 could drastically reduce the energy consumption of CO2 capture. Nat. Commun. 2018,
9, 2672.
26. Shi, H.C.; Zheng, L.N.; Huang, M.; Zuo, Y.H.; Kang, S.F.; Huang, Y.D.; Idem, R.; Tontiwachwuthikul, P.
Catalytic-CO2 -Desorption Studies of DEA and DEA-MEA Blended Solutions with the Aid of Lewis and
Bronsted Acids. Ind. Eng. Chem. Res. 2018, 57, 11505–11516. [CrossRef]
27. Teerawat Sema, A.N. Zhiwu Liang, Huancong Shi, Aravind V Rayer, Kazi Z Sumon, Pathamaporn
Wattanaphan, Amr Henni, Raphael Idem, Chintana Saiwan, Paitoon Tontiwachwuthikul, Part 5b: Solvent
chemistry: Reaction kinetics of CO2 absorptoin into reactive amine solutions. Carbon Manag. 2012, 3, 201–220.
28. Caplow, M. Kinetics of carbamate formation and breakdown. J. Am. Chem. Soc. 1968, 90, 6795–6803. [CrossRef]
29. Aboudheir, A.; Tontiwachwuthikul, P.; Chakma, A.; Idem, R. Kinetics of the reactive absorption of carbon
dioxide in high CO2 -loaded, concentrated aqueous monoethanolamine solutions. Chem. Eng. Sci. 2003, 58,
5195–5210. [CrossRef]
30. Donaldson, T.L.; Nguyen, N.Y. Carbon dioxide reaction kinetics and transport in aqueous amine membranes.
Ind. Eng. Chem. Fundam. 1980, 19, 260–266. [CrossRef]
31. Yu, W.C.; Astarita, G.; Savage, D.W. Kinetics of carbon dioxide absorption in solutions of
methyldiethanolamine. Chem. Eng. Sci. 1985, 40, 1585–1590. [CrossRef]
32. Shi, H.C.; Naami, A.; Idem, R.; Tontiwachwuthikul, P. 1D NMR Analysis of a Quaternary
MEA-DEAB-CO2 -H2 O Amine System: Liquid Phase Speciation and Vapor-Liquid Equilibria at CO2
Absorption and Solvent Regeneration Conditions. Ind. Eng. Chem. Res. 2014, 53, 8577–8591. [CrossRef]
33. Narku-Tetteh, J.; Muchan, P.; Saiwan, C.; Supap, T.; Idem, R. Selection of components for formulation of
amine blends for post combustion CO2 capture based on the side chain structure of primary, secondary and
tertiary amines. Chem. Eng. Sci. 2017, 170, 542–560. [CrossRef]
34. Muchan, P.; Saiwan, C.; Narku-Tetteh, J.; Idem, R.; Supap, T.; Tontiwachwuthikul, P. Screening tests of
aqueous alkanolamine solutions based on primary, secondary, and tertiary structure for blended aqueous
amine solution selection in post combustion CO2 capture. Chem. Eng. Sci. 2017, 170, 574–582. [CrossRef]
35. Singto, S.; Supap, T.; Idem, R.; Tontiwachwuthikul, P.; Tantayanon, S.; Al-Marri, M.J.; Benamor, A. Synthesis
of new amines for enhanced carbon dioxide (CO2 ) capture performance: The effect of chemical structure on
equilibrium solubility, cyclic capacity, kinetics of absorption and regeneration, and heats of absorption and
regeneration. Sep. Purif. Technol. 2016, 167, 97–107. [CrossRef]
36. Horwitz, W. Association of Official Analytical Chemists (AOAC) Methods; George Banta Co.: Menasha, WI,
USA, 1975.
37. Shi, H.C.; Zheng, L.N.; Huang, M.; Zuo, Y.H.; Li, M.Y.; Jiang, L.H.; Idem, R.; Tontiwachwuthikul, P. CO2
desorption tests of blended monoethanolamine-diethanolamine solutions to discover novel energy efficient
solvents. Asia-Pac. J. Chem. Eng. 2018, 13, e2186–e2199. [CrossRef]
38. Uyan, M.; Sieder, G.; Ingram, T.; Held, C. Predicting CO2 solubility in aqueous N-methyldiethanolamine
solutions with ePC-SAFT. Fluid Phase Equilib. 2015, 393, 91–100. [CrossRef]
39. Wangler, A.; Sieder, G.; Ingram, T.; Heilig, M.; Held, C. Prediction of CO2 and H2 S solubility and enthalpy of
absorption in reacting N-methyldiethanolamine /water systems with ePC-SAFT. Fluid Phase Equilib. 2018,
461, 15–27. [CrossRef]
40. Afari, D.B.; Coker, J.; Narku-Tetteh, J.; Idem, R. Comparative kinetic studies of solid absorber catalyst
(K/MgO) and solid desorber catalyst (HZSM-5)-aided CO2 absorption and desorption from aqueous
solutions of MEA and blended solutions of BEA-AMP and MEA-MDEA. Ind. Eng. Chem. Res. 2018, 57,
15824–15839. [CrossRef]

Sample Availability: Samples of the compounds are available from the authors, the solid catalysts are
commercially available, and the amines are also commercial available to prepare solutions.

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like