You are on page 1of 9

International Journal of Greenhouse Gas Control 86 (2019) 34–42

Contents lists available at ScienceDirect

International Journal of Greenhouse Gas Control


journal homepage: www.elsevier.com/locate/ijggc

Novel non-aqueous MEA solutions for CO2 capture T


Francis Bougie, Daniel Pokras, Xianfeng Fan

Institute for Materials and Processes, School of Engineering, University of Edinburgh, Mayfield Road, Edinburgh EH9 3JL, United Kingdom

ARTICLE INFO ABSTRACT

Keywords: Present obstacles for CO2 capture involve reducing capital and energy costs. This study describes carbon dioxide
Monoethanolamine capture with monoethanolamine (MEA) in organic solvents as a way to reduce the energy consumption during
Non-aqueous solutions desorption. The solvents used in this study are a mixture of ethylene glycol and 1-propanol (EG/PrOH), die-
CO2capture thylene glycol monoethyl ether (DEGMEE), and N-methylformamide (NMF). A microwave regeneration tech-
Energy efficiency
nique was used to easily and quickly screen the performances of the amine solutions and experiments were
Microwave regeneration
designed to investigate the solvent effects on CO2 capture by MEA. Globally, DEGMEE solutions resulted in the
lowest energy consumptions and good CO2 cyclic capacities; the best solutions reducing the energy consumption
by 78% in comparison to the traditional 30 wt% MEA aqueous solution. These findings show that organic sol-
vents have much promise in improving CO2 capture technology. It was also found that using NMF greatly
enhanced the CO2 absorption kinetics but was detrimental during the desorption step.

1. Introduction or piperazine (PZ), each possessing their own advantages in terms of


kinetics, CO2 capacity, degradation resistance or energy requirements
There is a worldwide concern over CO2 emission from our current (Bougie and Iliuta, 2010; Closmann et al., 2009; Gjernes et al., 2013;
systems of industrial and energy production and its effect on global Hartono et al., 2014; Martin et al., 2012; Rochelle et al., 2011). Water is
climate changes. This puts into question the available options govern- the chief solvent used in these solutions mainly due to its availability and
ments have access to implement limits on CO2 emissions. A popular low price. Unfortunately, water may not be the optimal solvent to use
option, which is receiving more investment each year, is the post- due, in part, to its high heat capacity, heat of vaporization and vapor
combustion Carbon Capture and Sequestration technology (CCS) (Lee pressure (Lail et al., 2014) which increase the energy requirements
et al., 2005; Price, 2014) to reduce CO2 emissions to sustainable levels during the regeneration step (El Hadri et al., 2017). Another variable to
while mitigating climate change effects. consider when using water as solvent is its detrimental ability to induce
At present, the most matured technology for CO2 capture is the post- degradation/corrosion of the amine and process equipment (Park et al.,
combustion flue gas absorption due to its ease of operation and scal- 2006; Zhao et al., 2011).
ability. This process is usually performed by aqueous alkanolamine so- With the aim of retaining the advantages of alkanolamine absorp-
lutions and monoethanolamine (MEA)-based absorbents are the most tion, while reducing the disadvantages of using water, a strategy to
popularly used in industry because of their low cost and high reactivity reduce the energy penalty of the regeneration step would be to use non-
with CO2 (Li et al., 2016; Lin et al., 2011; Luis, 2016; Oko et al., 2017). aqueous solutions. Such solutions have been proposed in the literature,
However, the regeneration of the solutions is the largest obstacle in replacing water with room temperature ionic liquids, CO2-binding li-
complete widespread implementation of CCS due to its high energy quids or organic solvents such as methanol or ethanol (Barzagli et al.,
penalty (House et al., 2009). Current researches mainly focus on im- 2014, 2013; Cui et al., 2018; Hart et al., 2010; Hasib-ur-Rahman and
proving the efficiency of this part of the process with refined techniques Larachi, 2012; Lail et al., 2014; Mathias et al., 2015; Park et al., 2006;
or development of new and more efficient amine solutions (Bougie and Tamajón et al., 2016; Tao et al., 2017; Wang et al., 2017; Zheng et al.,
Iliuta, 2012; Novotny et al., 2017; Patiño-Echeverri and Hoppock, 2012). 2014). Studies using non-aqueous solutions mentioned some direct
Recent papers have exhibited development of aqueous amine solutions advantages in CO2 capture as their fast kinetics or higher CO2 solubility
for CO2 capture with various amines such as diethanolamine (DEA), (Barzagli et al., 2013; Park et al., 2006; Tamajón et al., 2016) but very
triethanolamine (TEA), 2-amino-2-methyl-1-propanol (AMP), 2-amino-2- few of them measured the energy consumption and compared it with
hydroxymethyl-1,3-propanediol (AHPD), methyldiethanolamine (MDEA) their corresponding aqueous solutions. In this work, the focus is


Corresponding author.
E-mail address: X.Fan@ed.ac.uk (X. Fan).

https://doi.org/10.1016/j.ijggc.2019.04.013
Received 7 January 2019; Received in revised form 15 April 2019; Accepted 18 April 2019
Available online 30 April 2019
1750-5836/ © 2019 Elsevier Ltd. All rights reserved.
F. Bougie, et al. International Journal of Greenhouse Gas Control 86 (2019) 34–42

oriented toward developing efficient non-aqueous amine solutions a cold trap to avoid significant solvent loss, a non-dispersive infrared
based on their absorption and regeneration performances but also CO2 sensor (COZIR-W-100, CO2 calibration range: 0–70% v/v) and fi-
taking into account their regeneration energy consumption. nally a flow meter (Brooks Instruments SLA5860, 0–500 ml/min) before
Several properties need to be considered when selecting potential exiting through an exhaust line. The outlet CO2 flow rate was calculated
solvents, and low values of its heat capacity, vapor pressure and visc- by multiplying the CO2 sensor and flow meter readings. A blank ab-
osity are expected. The solvent also need to be available at low cost and sorption run was performed with an empty reactor to allow the ab-
it should show good solubility to the amine and CO2 products. Three sorbed quantity of CO2 to be calculated by integrating the difference
different solvents were selected in this study: (i) N-methylformamide between the CO2 breakthrough curves of the blank and sample. The
(NMF), (ii) diethylene glycol monoethyl ether (DEGMEE), and (iii) an CO2 loading in solution was calculated as the ratio of the absorbed
ethylene glycol/1-propanol mixture (EG/PrOH). quantity of CO2 (mol) to the amount of amine (mol) in solution. All
NMF was introduced as it is a common solvent found in laboratories absorption experiments were performed at atmospheric pressure and
with a high boiling point and low heat capacity and there was sig- room temperature and to avoid unexpected CO2 absorption by the so-
nificant interest in examining the effect of a highly polar solvent (high lution, all gas lines were purged with N2 before and between experi-
dielectric constant) with an amide group on CO2 absorption and deso- mental runs.
rption performances. To the best of our knowledge, no other studies
were found in the literature testing NMF for CO2 capture and the effect 2.3. Regeneration setup and procedure
of various polarity values has not been studied. DEGMEE on the other
hand, is a low polar solvent (low dielectric constant) with a high boiling The regenerations of the CO2-loaded MEA solutions were performed
point and a good representative of ether molecules. Ethers were found using the same quartz reactor as in the absorption experiments but
to be good potential solvents for CO2 capture due, in part, to their low using microwave irradiations to heat the sample. This technique was
CO2 Henry’s law constant (Henni et al., 2005; Li et al., 2014). A mixture used for its ease of operation and its ability to quickly heat the solution
of EG/PrOH was the additional solvent used in this study. EG is a sol- to the desired temperature while allowing the determination of the
vent commonly found in literature regarding CO2 capture (Barzagli energy consumption. The microwaves were produced by a magnetron
et al., 2014, 2013; Park et al., 2005; Xu et al., 2002) and the mixture operating at 2.45 GHz and controlled by an Alter SM445 power supply
with propanol produced a solvent containing many hydroxyl groups. with a maximum output power of 1.2 kW. A single mode waveguide
EG was mixed with propanol to reduce the viscosity of the solution directed the microwave energy towards respectively: (i) a dual-direc-
while maintaining a high boiling point as proposed by Barzagli tional coupler (GAE Inc., GA310x, calibrated to take into account the
(Barzagli et al., 2014). presence of the quartz reactor) which measure the forward and re-
Various MEA-based solutions with these solvents were tested flected microwave power flow, (ii) a resonant cavity where the liquid
through single CO2 absorption and regeneration steps as well as from sample is located, and (iii) a sliding short circuit (Sairem) to reflect the
cyclic experiments to assess their CO2 absorption and desorption per- microwave at normal incidence. Non-absorbed microwave power was
formances. Their energy consumptions were monitored during the absorbed by a water dummy-load. The temperature of the solution was
desorption steps to find which solutions required less energy while measured at the center of the liquid bulk every second with a fiber optic
having a high CO2 cyclic capacity. Results were also compared with sensor probe (Opsens OTG-MPK8).
those of the well-known 30 wt% MEA aqueous solution. To the best of Prior to the start of the microwave irradiation, a nitrogen flow was
our knowledge, similar results are not available in the open literature established at 100 ml/min to purge the gas lines from possible residual
but are essential for further improvement in the design of energy-effi- CO2 and then acted as a sweeping gas. The microwave source was
cient CO2 absorbent solutions. turned on at an initial power of 125 W to heat the solution from am-
bient to the desired regeneration temperature and then the power was
2. Methods & experimental manually reduced and controlled to maintain a constant temperature.
The installed dual-directional couplers measured the forward and re-
2.1. Reagents flected microwave power flows and the difference between them re-
presented the amount of energy absorbed by the solution. At the end of
In this work, all solutions were prepared using MEA (CAS No. 141- the regeneration time, the magnetron was turned off and the solution
43-5, 98%, Sigma-Aldrich UK) and DEGMEE (CAS No. 111-90-0, 99%, was allowed to cool down. The amount of stripped CO2 was determined
Honeywell), NMF (CAS No. 123-39-7, 99%, Acros Organics), or an EG by direct integration of the outlet CO2 flow rate. For all the tested so-
(CAS No. 107-21-1, 99%, Sigma-Aldrich UK) and PrOH (CAS No. 71-23- lution, it was visually observed that the solution quantity was qualita-
8, 99%, Acros Organics) mixture (1:1 vol ratio at 25 °C). A Sartorius tively at the same level than before the regeneration step.
ED224S balance with a precision of ± 1 × 10−4 g was used to prepare
the solutions and it was calculated that the uncertainty of the reported 2.4. Studied systems
concentrations was less than 0.1 wt%. Gases used for absorption and
desorption experiments (CO2 and N2) were of commercial grade with a As the first test, MEA solutions (20, 30, 40, 50, 60 and 70 wt%) with
minimum purity of 99.99% (Linde Group UK). each of the selected solvent (DEGMEE, NMF and EG/PrOH) were
characterized by a single CO2 absorption and desorption step. The se-
2.2. Absorption setup and procedure lected range of concentration is based on our previous work (Bougie
and Fan, 2018) which showed that the MEA concentration needs to be
As an exhaustive description of the absorption/regeneration setup high enough to sustain a fast absorption kinetic but not too high to
can be found in our previous work (Bougie and Fan, 2018), only the avoid highly viscous solutions. For these experiments, the CO2 ab-
important characteristics will be summarized here. All experiments sorption was accomplished until the solution was almost saturated
started with an absorption step where 5 g of solution was contained in a which took between 30 and 50 min depending on the MEA concentra-
small cylindrical quartz reactor which had an inside diameter of 14 mm tion. Then the solutions were regenerated at 90 °C for 50 min. Based on
and a wall thickness of 1.5 mm. A binary gas mixture (20% CO2 and the absorption capacity and regeneration performances (quantity of
80% N2) was bubbled at the bottom of the solution through a small stripped CO2 and energy consumption), the best solutions were sub-
quartz tube at a total flow rate of 100 ml/min via two calibrated mass mitted to cyclic absorption-regeneration experiments to mimic the in-
flow controllers (Brooks Instruments GF-Series, 0–400 ml/min N2, dustrial CO2 capture process. Absorption steps of 15 min were followed
0–100 ml/min CO2). The outlet gas stream passed successively through by 12 min of regeneration at 90 °C. Several cycles were done until the

35
F. Bougie, et al. International Journal of Greenhouse Gas Control 86 (2019) 34–42

Fig. 1. Values of various physico-chemical properties for selected pure organic solvents and MEA in comparison to water; (a) viscosity at 20 °C; (b) vapor pressure at
25 °C; (c) specific heat capacity at 25 °C; and (d) dipole moment. *Estimated values based on mole fraction of each component in the mixture.

absorbed CO2 quantity was similar to the stripped one, indicating that a circumvent this problem, EG was mixed in a 1:1 vol ratio with propanol
steady-state condition was reached. This usually took up to 4 cycles (μ = 2.26 mPa·s at 20 °C (Kroschwitz and Seidel, 2004)) to reduce the
depending on the MEA concentration and nature of the solvent. viscosity while maintaining a high boiling point (˜150 °C) as proposed
by Barzagli (Barzagli et al., 2014). Propanol was used instead of the
3. Results and discussion traditional methanol or ethanol for its higher boiling point and lower
evaporation tendency while higher alcohols (e.g. butanol, pentanol,
3.1. Solvent properties hexanol) were discarded to avoid solubility issues with the amines and
products obtained after CO2 absorption.
MEA was selected for this study as it is considered as a benchmark
amine for CO2 capture in the literature (Luis, 2016). The solvents NMF, 3.2. CO2 absorption experiments
DEGMEE and EG/PrOH were chosen for their low heat capacity, rela-
tively high boiling point (so low vapor pressure), and low cost (Bipp As explained in Section 2.4., CO2 absorption experiments were first
and Kieczka, 2011; Chiu and Li, 1999; Daubert, 1989; Dow Chemical performed with all the selected solutions and their absorption capacity
Company, 1989; Lewis, 2007; Riddick et al., 1985). Some important profiles are presented in Fig. 2 to visualize the solvent’s effects on the
physico-chemical properties of these solvents (Li et al., 2009), MEA loading capacity and kinetics of CO2 absorption by MEA.
(Mandal et al., 2003) and water are shown in Fig. 1. Estimated prop- From Fig. 2 (a–c), it is possible to see, as expected, that the total
erties of the EG/PrOH solvent were also included, based on mole amount of absorbed CO2 increased with the amine concentration and
fraction of each component in the mixture. They were included for that generally, the absorbed quantity of CO2 at a specified MEA con-
qualitative comparisons and discussions only as no precise calculations centration (as in Fig. 2d) at the end of the absorption time follow this
were made with these data. It should be noted that all data included in trend: water > NMF > DEGMEE > EG/PrOH. The difference in the
figures in this work can be found in a dataset available online (Bougie final CO2 loading (e.g. highest at 0.50 mol/mol in water and lowest at
et al., 2019). 0.47 mol/mol in EG/PrOH for the 30 wt% MEA solutions) is however
As mentioned in the Introduction section, NMF (CH3-NH−CHO) quite small indicating a limited effect of the solvent on the total ab-
was introduced into the study as there was significant interest in ex- sorption capacity. This shows that the majority of the CO2 reacted
amining the effect of a highly polar solvent with an amide group in chemically with the amine and that the CO2 partial pressure was not
combination with MEA on CO2 absorption and desorption. To the best high enough to modify significantly the amount of physically dissolved
of our knowledge, no other studies were found in the literature using CO2 in solutions. The presence of bicarbonate/carbonate in the aqueous
NMF as a solvent for CO2 capture. DEGMEE (C2H5-O−CH2CH2- solution may explain the slightly higher absorption capacity using
O−CH2CH2OH) on the other hand, is a low polar solvent and a good water (Dang and Rochelle, 2003; Idris et al., 2014).
representative of an ether molecule. Ethers were found to be good po- Results displayed in Fig. 2(a, b, d) show that the kinetics of CO2
tential solvents for CO2 capture due, in part, to their low CO2 Henry’s absorption by MEA in EG/PrOH and DEGMEE are similar and that more
law constant (Henni et al., 2005; Li et al., 2014). A mixture of EG/PrOH concentrated solutions have better kinetics; the initial slope of their
(HO−CH2CH2−OH/CH3CH2CH2−OH) was the additional solvent absorption curves are higher. The CO2 absorption is clearly different in
used in this study. EG is a solvent commonly found in literature re- NMF (Fig. 2c) where all the curves are overlapping initially and then
garding CO2 capture (Barzagli et al., 2014, 2013; Park et al., 2005; Xu diverged one by one instead of spreading in an array of curves like in
et al., 2002) and the mixture with propanol produced a solvent con- Fig. 2a or b. A more spectacular way to notice the difference in ab-
taining many hydroxyl groups. EG was not used as a sole solvent due to sorption behaviour in NMF is to plot the CO2 absorption rate, as pre-
its high viscosity (μ = 19.9 mPa·s at 20 °C (Barzagli et al., 2014)) which sented in Fig. 3a for all the tested solutions in NMF and in Fig. 3(b–c)
could lower the solutions mass transfer and absorption efficiency. To where comparisons of the 30 wt% MEA solutions are shown. On Fig. 3b,

36
F. Bougie, et al. International Journal of Greenhouse Gas Control 86 (2019) 34–42

Fig. 3. CO2 absorption rate for MEA solutions in (a) NMF. In (b), curves for
30 wt% MEA solutions in all tested solvents as a function of time and in (c), as a
function of the CO2 loading (mol/mol).

it is possible to see that the non-aqueous MEA solution with DEGMEE


(and EG/PrOH) or even the aqueous one absorbed up to approximately
90% of the 20 ml/min CO2 bubbled through the solution for only 2 min
in most cases and continued to absorb but progressively less efficiently
afterward. However, with NMF, almost all of the CO2 was absorbed for
approximately 11 min indicating that for NMF-containing solutions, the
absorption was limited by the supplied amount of CO2 and that the
absorption kinetics was greatly improved in comparison to the other
Fig. 2. CO2 absorption profiles for various MEA concentrations in (a) EG/PrOH; solvents. In Fig. 3b, the 30 wt% MEA in NMF was more efficient for the
(b) DEGMEE; and (c) NMF. Comparison in (d) of the 30 wt% MEA solutions first 14 min at which point the CO2 loading (0.42) was 17% and 27%
with the tested solvents and with water. Achieved loading (mol/mol) are in- greater than that in MEA + water (0.35) and MEA + EG/PrOH or
dicated in bold on each curves.
DEGMEE (0.32), respectively. Fig. 3c shows as expected that whatever
the solvent used for absorption, the absorption rate for CO2 tends

37
F. Bougie, et al. International Journal of Greenhouse Gas Control 86 (2019) 34–42

toward zero as a loading near 0.5 is reached which indicate that almost
all the free MEA in these solutions have reacted.

3.2.1. CO2 absorption in the MEA-NMF solutions


A possible explanation to an increased absorption rate in the NMF-
containing solutions could be linked to the molecular structure of NMF
(CH3-NH−CHO) where the presence of an amide functional group may
have a significant effect on the CO2 absorption. Pure NMF was tested for
its CO2 absorption capacity but as no significant absorption occurred,
an important physical CO2 solubility or a direct reaction of NMF with
CO2 (chemical solubility) had to be discarded.
Current speculations for the improvement in absorption rates for the
amine-amide solutions may come from a modification of the pKa of
MEA in NMF. Cox (Cox, 2013) reported various amines having a higher
pKa in formamide than in water (e.g. an increase of 11% of the pKa of
aniline) so a similar phenomenon may happen here in NMF. A higher Fig. 4. Maximum temperature achieved during the CO2 absorption in various
pKa for MEA would result in a higher reaction rate constant k2 (m3/ MEA solutions. Lines were included to show the trends.
kmol.s, see Eq. 1) as predicted by the analysis of Brønsted plots for
many amines used for CO2 capture (Bougie and Iliuta, 2009; Derks would be counterbalanced by the centrifugal force of the rotating de-
et al., 2006). Furthermore, taking into consideration the zwitterion vice which enhances the mass transfer coefficients by at least one order
mechanism (Eqs. 1 and 2), a widely accepted mechanism used to de- of magnitude (Chen et al., 2005; Wang et al., 2015; Zhang et al., 2011).
scribe the CO2 absorption by amines such as MEA (Lv et al., 2015), a
faster deprotonation step of the MEA zwitterion (kB) would improve the
3.2.3. Heat generation during absorption
kinetics. In the literature, several amines, as well as MEA, were found to
The CO2 absorption by an amine solution is an exothermal process
have an apparent reaction order a little bit higher than 1 due to the fact
and the heat generated within the solution can be an important factor to
that the deprotonation step is not totally instantaneous (Usubharatana
take into account. The absorptions were not carried out isothermally to
and Tontiwachwuthikul, 2009). Therefore, bases in solution (such as a
investigate this phenomenon (as it will occur industrially). The rise in
stronger MEA) can deprotonate the zwitterion (Eq. 2) to form a car-
temperature can influence the solubility, the kinetics and the design of
bamate. NMF, with its amide group and high dielectric constant, can
the absorber which can, for example, include intercooling sections
also act as a proton acceptor or be involved in this step, accelerating it
(Rezazadeh et al., 2017). Fig. 4 displays the maximum temperatures
(Cox, 2013).
achieved during the CO2 absorption process by the tested MEA solu-
k2 tions before the solutions returned gradually to room temperature as
CO2 + RNH2 RNH2+COO (1)
the absorption steps were long enough (between 30 and 50 min de-
kB pending on the MEA concentration).
RNH2+COO + Base RNHCOO + BaseH+ (2) Clearly, for all concentrations, the maximum temperatures follow
the ranking: NMF > DEGMEE > EG/PrOH > water. This result is
quite surprising considering that the specific heat capacities for the
3.2.2. Effect of CO2 absorption on solvent viscosity
organic solvents are similar. For NMF, higher temperatures were
During the absorption process, it was noticed that the MEA solutions
achieved due to the faster kinetics; more CO2 was absorbed, releasing
became more and more viscous as the MEA concentration or the CO2
more energy. For water, its higher heat capacity explains the lowest
loading increased. High viscosity values are often achieved during CO2
temperatures. It is believed that EG/PrOH solutions achieved lower
absorption by researchers using non-aqueous solutions (Mathias et al.,
maximum temperatures than that of DEGMEE during absorption due to
2015). For comparison purpose and to analyze the behaviour of the
the volatility of propanol as well as the higher heat capacity of EG in
different solutions, viscosity values were measured before and after the
comparison to DEGMEE. One fraction of the heat produced by the re-
CO2 absorption at 25 °C for the 30 wt% MEA solutions and results are
action was used to vaporize propanol (which condensed and returned to
reported in Table 1. It can be seen that the higher the viscosity of the
the reactor), reducing the temperature in those systems.
pure solvent (Fig. 1), the higher the viscosity of the fresh solutions.
However, for the loaded solutions, the DEGMEE solution surprisingly
achieved a higher viscosity than that of EG/PrOH mixture, leading to a 3.3. CO2 desorption
cloudy and opaque solution. High viscosities could be a detrimental
issue for solutions used in the CO2 capture process, but recent studies Following the absorption step, an important part of this study was to
have shown that the impact of a high viscosity is not so severe and that analyze the regeneration performance of the tested solutions. As men-
these solutions can be used when using high performance absorbers or tioned in Sections 2.3 and 2.4, the regenerations were performed at
desorbers like a rotating packed bed contactor (RPB). In a RPB, the 90 °C for 50 min (except for NMF solutions as it will be explained later)
reduction of the mass transfer when using highly viscous solutions and the obtained desorption profiles are presented in Fig. 5 while Fig. 6
shows the total amount of CO2 desorbed from the tested solutions.
Table 1 Globally, it can be seen from Figs. 5(a–c) that the CO2 desorption
Viscosity values of 30 wt% MEA solutions before and after CO2 absorption at pattern for the solvents used are very similar. Solutions with low MEA
25 °C. concentrations (20 and 30 wt%) have a high initial desorption rate but
Solvent Viscosity (mPa·s) the quantity of the final released CO2 is not as high as that from the 40
and 50 wt% solutions due to their lower absolute amounts of dissolved
fresh solution loaded solution CO2 during the absorption step. When MEA concentration is higher
EG/PrOH 12.8 63.7
than 40 wt%, the desorption rates and the amount of CO2 stripped
DEGMEE 9.4 98.5
NMF0 3.1 22.9 during the regeneration period gradually decreased with an increase of
Water 2.5 3.7 the MEA concentration, possibly due to a corresponding increase of the
solution viscosity. Even at the regeneration temperature, the high

38
F. Bougie, et al. International Journal of Greenhouse Gas Control 86 (2019) 34–42

Fig. 6. CO2 desorbed after the regeneration step.

Figs. 5d and 6 show that the quantity of CO2 released during the
desorption process was generally higher from EG/PrOH or DEGMEE
non-aqueous solutions than from aqueous solutions. This corresponds to
a reverse trend from what is seen in Fig. 3b concerning the absorption
rate and show that a solvent improving the absorption will generally
have the opposite effect during desorption. Similarly, the amount of
CO2 stripped from MEA-NMF is much smaller than from the other
solvents. The reason for this poor performance may be the same used to
explain the fast absorption kinetics of CO2 with NMF in Section 3.2.1.: if
MEA is a stronger base (higher pKa) in NMF, the bonds with the ab-
sorbed CO2 in the carbamate (RNHCO2−) will be more difficult to break
leading to reduced desorption capacities. It was also noticed that during
the regeneration of the rich MEA-NMF solutions, some crystallization
occurred with crystals forming on the quartz reactor wall just above the
meniscus of the solution. This is the reason why these regenerations
were stopped earlier than with the other solvents.

3.3.1. Energy consideration


Most of current research on reversible CO2 capture centers either on
CO2 capture or regeneration performance of various solutions but rarely
assess the energy consumption of the regeneration step. In this work, as
explained in Section 2.3, the dual-directional couplers installed on the
microwave setup allowed to determine the quantity of energy absorbed
by the solution during the regeneration steps. It should be mentioned
that the desorption energies measured are specific for the setup used in
this study and can hardly be compared to values from literature. It is
known that the microwave desorption energy measured for small
samples (as 5 g in this work) overestimate the real energy consumption
to strip the CO2 (Bermúdez et al., 2015) due, for example, to a high
surface/volume ratio of the sample resulting in relatively high heat
losses or evaporation rate of the solutions. The heat taken out by the N2
purging gas and the heat consumed for heating the quartz reactor also
interfered. All these energies were not consumed for CO2 desorption,
but are included in the reported data. However, as all experiments are
performed under the same conditions, the energy consumptions can be
used here to compare different solutions and analyze their regeneration
performances: a solution consuming more energy in this work will si-
milarly consume more energy industrially.
To compare the regeneration performances of the solutions, a metric
was elaborated. During regeneration, an ideal solution will display a
Fig. 5. CO2 desorption profiles for various MEA concentrations in (a) EG/PrOH; low energy/released CO2 ratio EC (Energy per CO2 in kJ/mol) as de-
(b) DEGMEE; and (c) NMF. Comparison in (d) of the 30 wt% MEA solutions fined by Eq. (3). This factor was calculated every two minutes during
with the tested solvents and with water. Final residual loading (mol/mol) after regeneration and represents the cumulative energy consumed until a
desorption are indicated in bold on each figure. given time divided by the cumulative quantity of CO2 stripped before
the same time. Results for all tested solutions are included in Fig. 7.
viscosity of the 60 and 70 wt% MEA EG/PrOH solutions was noticeable
Energy (kJ )
experimentally (and in Fig. 6) reducing the released quantity of CO2 to EC =
CO2 stripped (mol) (3)
values lower than with DEGMEE or water.

39
F. Bougie, et al. International Journal of Greenhouse Gas Control 86 (2019) 34–42

Clearly, there is a significant influence of the solvent used on the EC


values obtained. With EG/PrOH, the minimum values at low re-
generation times are around 1200 kJ/mol for the 20–40 wt% MEA so-
lutions while they can be as low as around 400 kJ/mol with DEGMEE
(20–30 wt% MEA) or around 900 kJ/mol with NMF (30–40 wt% MEA).
A comparative test using the conventional 30 wt% MEA solution shows
in Fig.7d that all non-aqueous solutions (in particular the one with
DEGMEE) performed better than the aqueous one whose minimum EC
values are around 3000 kJ/mol. The use of non-aqueous solutions is
then obviously reducing the energy consumption during the regenera-
tion process thanks in large part to the low heat capacity of the organic
solvents (see Fig. 1c).
From Fig.7, it is also possible to see that usually the EC values follow
a concave curve with an optimum (minimum) that shift depending on
the MEA concentration. In the first few minutes of desorption, EC is
high because the initial energy consumption is used to increase the
temperature of the solution with few released CO2. Then, a lower en-
ergy consumption and high quantity of stripped CO2 reduce EC to a
minimum. Finally, as energy is continuously supplied to the solution
with less and less CO2 being stripped, the EC ratio starts to rise again.
There is, however, one exception to this trend from the lowest MEA-
concentrated DEGMEE solutions (20 to 40 wt% MEA) where the EC
curves start at low values and steadily increased (Fig. 7b). It is believed
that this beneficial phenomenon (as low values of EC are achieved)
happened due to the use of microwaves to regenerate the solutions.
Theoretically, microwave heating is based on the ability of molecules
with a dipole moment to absorb microwave energy and effectively
convert it into heat. The dielectric constant of a solvent is involved and
is a measure of the substance’s ability to absorb and store microwave
energy (Cherbański and Molga, 2009). Among the tested solvents,
DEGMEE has the lowest dielectric constant (Gordon and Ford, 1972)
(12.6 for DEGMEE, 37.7 for EG, 78.5 for water, and 182 for NMF at
25 °C) so it is possible that the microwaves interacted preferentially
with the MEA carbamates in DEGMEE solutions, especially those with
high DEGMEE content (low MEA concentrations) to desorb more effi-
ciently the CO2, which would result in low EC values. To validate the
presence of this beneficial non-thermal microwave effect, a thorough
study including a comparison of microwave heating to a conventional
heating mode performed under the same conditions is needed and is
currently in progress in our laboratory.

3.3.2. Cyclic absorption and regeneration process


Based on results from single absorption and regeneration steps,
some solutions were selected to perform cyclic experiments as ex-
plained in Section 2.4 in order to emulate the industrial steady state
CO2 capture process. These solutions were 40 wt% MEA in EG/PrOH
(representative of EG/PrOH solutions with low EC values), 20 to 40 wt
% MEA in DEGMEE (representatives of DEGMEE solutions with low EC
values) and 30 and 50 wt% MEA in water (representatives of aqueous
solutions including the conventional 30 wt% MEA solution). No NMF
solutions were selected due to the crystallization issue during re-
generation. Several cycles were done with these solutions until the
absorbed CO2 quantity was similar to the stripped one, indicating that a
steady-state condition was reached. This usually took 3 or 4 cycles
depending on the MEA concentration and nature of the solvent.
Steady-state results from these experiments are displayed in Table 2
where it is possible to see that all non-aqueous solutions performed
better than the aqueous solutions comparing their EC values. DEGMEE
solutions have the lowest EC values and, as observed in Fig. 7b, the
20 wt% MEA is the best, having an EC value of only 785 kJ/mol. This
Fig. 7. EC results for various MEA solutions with (a) EG/PrOH; (b) DEGMEE; represents a reduction of 78% of the energy requirement per mol of CO2
(c) NMF. Inserts covering the first 10-minute period are also included. in comparison to the well-known 30 wt% MEA aqueous solution. The
Comparison in (d) of the 30 wt% MEA solutions with the tested solvents and aqueous 50 wt% MEA solution also performed better than the 30 wt%
with water. aqueous solution with a reduction of 10% of its EC value. This confirm
the finding of our previous work (Bougie and Fan, 2018) stating that
among the aqueous MEA solutions, the 50 wt% one performed better.

40
F. Bougie, et al. International Journal of Greenhouse Gas Control 86 (2019) 34–42

Table 2 Chemistry.
Results of the cycling absorption-regeneration process at steady-state for var- Bougie, F., Fan, X., 2018. Microwave regeneration of monoethanolamine aqueous solu-
ious MEA solutions. tions used for CO2 capture. Int. J. Greenhouse Gas Control 79, 165–172.
Bougie, F., Iliuta, M.C., 2009. Kinetics of absorption of carbon dioxide into aqueous so-
Solvent MEA wt% CO2 cyclic capacity (mol) Energy (J) EC (kJ/mol) lutions of 2-amino-2-hydroxymethyl-1,3-propanediol. Chem. Eng. Sci. 64, 153–162.
Bougie, F., Iliuta, M.C., 2010. CO2 absorption into mixed aqueous solutions of 2-Amino-2-
EG/PrOH 40 0.0060 10193 1700 hydroxymethyl-1,3-propanediol and piperazine. Ind. Eng. Chem. Res. 49,
DEGMEE 20 0.0053 4140 785 1150–1159.
Bougie, F., Iliuta, M.C., 2012. Sterically hindered amine-based absorbents for the removal
30 0.0052 4890 929
of CO2 from gas streams. J. Chem. Eng. Data 57, 635–669.
40 0.0047 6018 1288
Bougie, F., Pokras, D., Fan, X., 2019. Article MEA Non-aqueous Data. Mendeley Data, v1.
Water 30 0.0044 16129 3630
https://doi.org/10.17632/t2r2c8sx9h.1.
50 0.0035 11524 3257 Chen, Y.-S., Lin, C.-C., Liu, H.-S., 2005. Mass transfer in a rotating packed bed with vis-
cous newtonian and non-newtonian fluids. Ind. Eng. Chem. Res. 44, 1043–1051.
Cherbański, R., Molga, E., 2009. Intensification of desorption processes by use of mi-
4. Conclusions crowaves—an overview of possible applications and industrial perspectives. Chem.
Eng. Process. Process. Intensif. 48, 48–58.
Chiu, L.F., Li, M.H., 1999. Heat capacity of alkanolamine aqueous solutions. J. Chem.
In this study, novel non-aqueous MEA solutions were studied in Eng. Data 44, 1396–1401.
order to potentially reduce the energy consumption during the CO2 Closmann, F., Nguyen, T., Rochelle, G.T., 2009. MDEA/Piperazine as a Solvent for CO2
Capture 1. pp. 1351–1357.
capture process. Three different solvents were selected: an ether (die- Cox, B.G., 2013. Acids and Bases : Solvent Effects on Acid-base Strength.
thylene glycol monoethyl ether - DEGMEE), an amide (N-methylfor- Cui, M., Chen, S., Qi, T., Zhang, Y., 2018. Investigation of CO2 capture in nonaqueous
mamide - NMF), and a hydroxyl-containing mixture (ethylene glycol/1- ethylethanolamine solution mixed with porous solids. J. Chem. Eng. Data 63,
1198–1205.
propanol -EG/PrOH). Various MEA solutions with these solvents were
Dang, H., Rochelle, G.T., 2003. CO2 absorption rate and solubility in Monoethanolamine/
tested through absorption and regeneration steps as well as from cyclic Piperazine/Water. Sep. Sci. Technol. 38, 337–357.
experiments to assess their CO2 absorption and desorption perfor- Daubert, T.E., 1989. Physical and thermodynamic properties of pure chemicals: data
compilation. In: Danner, R.P. (Ed.), Hemisphere Pub. Corp., New York.
mances. The energy consumptions were also monitored during the
Dow Chemical Company, 1989. The Glycol Ethers Handbook. Dow Chemical Company.
desorption steps to find solutions requiring less energy per mol of Derks, P.W.J., Kleingeld, T., van Aken, C., Hogendoom, J.A., Versteeg, G.F., 2006.
stripped CO2 than with the conventional 30 wt% MEA aqueous solu- Kinetics of absorption of Carbon Dioxide in aqueous piperazine solutions. Chem. Eng.
tion. Sci. 61, 6837–6854.
El Hadri, N., Quang, D.V., Goetheer, E.L.V., Abu Zahra, M.R.M., 2017. Aqueous amine
From the CO2 absorption point of view, NMF solutions showed a solution characterization for post-combustion CO2 capture process. Appl. Energy 185,
spectacular increase of the kinetics due to a possible increase of the pKa 1433–1449.
of MEA when dissolved in NMF and by the presence of the amide group Gjernes, E., Helgesen, L.I., Maree, Y., 2013. Health and environmental impact of amine
based post combustion CO2 capture. Energy Procedia 37, 735–742.
and high dielectric constant of NMF which may speed up the depro- Gordon, A.J., Ford, R.A., 1972. The Chemist’s Companion: a Handbook of Practical Data,
tonation step of the zwitterions into carbamates. However, bad deso- Techniques, and References. Wiley, New York.
rption performances and the formation of crystals on the wall of the Hart, R., Pollet, P., Hahne, D.J., John, E., Llopis-Mestre, V., Blasucci, V., Huttenhower, H.,
Leitner, W., Eckert, C.A., Liotta, C.L., 2010. Benign coupling of reactions and se-
reactor limited further use of this solvent. parations with reversible ionic liquids. Tetrahedron 66, 1082–1090.
The best results came from the use of the DEGMEE solutions, Hartono, A., Ciftja, A.F., Brúder, P., Svendsen, H.F., 2014. Characterization of amine-
especially those at low MEA concentration (20–40 wt%). These solu- impregnated adsorbent for CCS post combustion. Energy Procedia 63, 2138–2143.
Hasib-ur-Rahman, M., Larachi, F., 2012. CO2Capture in Alkanolamine-RTIL Blends via
tions showed good absorption and desorption rates but also low energy
Carbamate Crystallization: Route to Efficient Regeneration. Environ. Sci. Technol. 46,
consumptions during the regeneration steps. The EC value obtained for 11443–11450.
the 20 wt% MEA in DEGMEE during the cycling process was found to be Henni, A., Tontiwachwuthikul, P., Chakma, A., 2005. Solubilities of Carbon Dioxide in
polyethylene glycol ethers. Can. J. Chem. Eng. 83, 358–361.
78% lower than the value obtained for the conventional aqueous 30 wt
House, K.Z., Harvey, C.F., Aziz, M.J., Schrag, D.P., 2009. The energy penalty of post-
% MEA solution. These results are a significant first step for further combustion CO2 capture & storage and its implications for retrofitting the U.S. in-
improvement, for example by studying other potential ether molecules stalled base. Energy Environ. Sci. 2, 193–205.
in the future. The use of microwaves to regenerate the solutions com- Idris, Z., Jens, K.J., Eimer, D.A., 2014. Speciation of MEA-CO2 adducts at equilibrium
using raman spectroscopy. Energy Procedia 63, 1424–1431.
bined with the low dielectric constant of DEGMEE may explain these Kroschwitz, J.I., Seidel, A., 2004. Kirk-Othmer Encyclopedia of Chemical Technology.
favorable results; the microwaves coupling preferentially and re- Wiley-Interscience, Hoboken, N.J.
generating the MEA carbamates more efficiently than with the use of Lail, M., Tanthana, J., Coleman, L., 2014. Non-aqueous solvent (NAS) CO2 capture pro-
cess. Energy Procedia 63, 580–594.
the other solvents. However, additional research, which is in progress in Lee, A., Christensen, D., Cappelen, F., Hartog, J., Thompson, A., Johns, G., Senior, B.,
our laboratory, concerning the solvent’s dielectric constant influence Akhurst, M., 2005. Policies and incentives developments in CO2 capture and storage
through comparison with conventional heating are required to confirm technology: a focused survey by the CO2 capture project. Chapter 1 In: Thomas, D.C.
(Ed.), Carbon Dioxide Capture for Storage in Deep Geologic Formations. Elsevier
the presence of this beneficial effect. Results will be published in a soon Science, Amsterdam, pp. 17–35.
coming publication. Lewis Sr., R.J., 2007. Hawley’s Condensed Chemical Dictionary / Richard J. Lewis. Sr.
John Wiley & Sons, Hoboken, N.J.
Li, X., Xu, G., Wang, Y., Hu, Y., 2009. Density, viscosity, and excess properties for binary
Acknowledgements
mixture of diethylene glycol monoethyl ether + water from 293.15 to 333.15 K at
atmospheric pressure. Chin. J. Chem. Eng. 17, 1009–1013.
Authors gratefully acknowledge theEngineering and Physical Li, Y., Huang, W., Zheng, D., Mi, Y., Dong, L., 2014. Solubilities of CO2 capture absorbents
2-ethoxyethyl ether, 2-butoxyethyl acetate and 2-(2-ethoxyethoxy)ethyl acetate.
Sciences Research Council (EPSRC) of the UK [grant number EP/
Fluid Phase Equilibr. 370, 1–7.
N024672/1]. Li, K., Cousins, A., Yu, H., Feron, P., Tade, M., Luo, W., Chen, J., 2016. Systematic study of
aqueous monoethanolamine-based CO2 capture process: model development and
References process improvement. Energy Sci. Eng. 4, 23–39.
Lin, Y.-J., Pan, T.-H., Wong, D.S.-H., Jang, S.-S., Chi, Y.-W., Yeh, C.-H., 2011. Plantwide
control of CO2 capture by absorption and stripping using monoethanolamine solu-
Barzagli, F., Mani, F., Peruzzini, M., 2013. Efficient CO2 absorption and low temperature tion. Ind. Eng. Chem. Res. 50, 1338–1345.
desorption with non-aqueous solvents based on 2-amino-2-methyl-1-propanol (AMP). Luis, P., 2016. Use of monoethanolamine (MEA) for CO2 capture in a global scenario:
Int. J. Greenhouse Gas Control 16, 217–223. consequences and alternatives. Desalination 380, 93–99.
Barzagli, F., Lai, S., Mani, F., 2014. Novel non-aqueous amine solvents for reversible CO2 Lv, B., Guo, B., Zhou, Z., Jing, G., 2015. Mechanisms of CO2 capture into mono-
capture. Energy Procedia 63, 1795–1804. ethanolamine solution with different CO2 loading during the Absorption/Desorption
Bermúdez, J.M., Beneroso, D., Rey-Raap, N., Arenillas, A., Menéndez, J.A., 2015. Energy processes. Environ. Sci. Technol. 49, 10728–10735.
consumption estimation in the scaling-up of microwave heating processes. Chem. Mandal, B.P., Kundu, M., Bandyopadhyay, S.S., 2003. Density and Viscosity of Aqueous
Eng. Process. Process. Intensif. 95, 1–8. Solutions of (N-Methyldiethanolamine + Monoethanolamine), (N-
Bipp, H., Kieczka, H., 2011. Formamides, Ullmann’s Encyclopedia of Industrial Methyldiethanolamine + Diethanolamine), (2-Amino-2-methyl-1-propanol +

41
F. Bougie, et al. International Journal of Greenhouse Gas Control 86 (2019) 34–42

Monoethanolamine), and (2-Amino-2-methyl-1-propanol + Diethanolamine). J. Solvents, 4th eds. John Wiley and Sons, New York, NY.
Chem. Eng. Data 48, 703–707. Rochelle, G., Chen, E., Freeman, S., Van Wagener, D., Xu, Q., Voice, A., 2011. Aqueous
Martin, S., Lepaumier, H., Picq, D., Kittel, J., de Bruin, T., Faraj, A., Carrette, P.-L., 2012. piperazine as the new standard for CO2 capture technology. Chem. Eng. J. 171,
New amines for CO2 capture. IV. degradation, corrosion, and quantitative structure 725–733.
property relationship model. Ind. Eng. Chem. Res. 51, 6283–6289. Tamajón, F.J., Álvarez, E., Cerdeira, F., Gómez-Díaz, D., 2016. CO2 absorption into N-
Mathias, P.M., Zheng, F., Heldebrant, D.J., Zwoster, A., Whyatt, G., Freeman, C.M., methyldiethanolamine aqueous-organic solvents. Chem. Eng. J. 283, 1069–1080.
Bearden, M.D., Koech, P., 2015. Measuring the absorption rate of CO2 in nonaqueous Tao, M., Gao, J., Zhang, P., Zhang, W., Liu, Q., He, Y., Shi, Y., 2017. Biogas upgrading by
CO2-Binding organic liquid solvents with a wetted-wall apparatus. ChemSusChem 8, capturing CO2 in non-aqueous phase-changing diamine solutions. Energy Fuels 31,
3617–3625. 6298–6304.
Novotny, V., Vitvarova, M., Kolovratnik, M., Hrdina, Z., 2017. Minimizing the energy and Usubharatana, P., Tontiwachwuthikul, P., 2009. Enhancement factor and kinetics of CO2
economic penalty of CCS power plants through waste heat recovery systems. Energy capture by MEA-methanol hybrid solvents. Energy Procedia 1, 95–102.
Procedia 108, 10–17. Wang, M., Joel, A.S., Ramshaw, C., Eimer, D., Musa, N.M., 2015. Process intensification
Oko, E., Wang, M., Joel, A.S., 2017. Current status and future development of solvent- for post-combustion CO2 capture with chemical absorption: a critical review. Appl.
based carbon capture. Int. J. Coal Sci. Technol. 4, 5–14. Energy 158, 275–291.
Park, S.W., Lee, J.W., Choi, B.S., Lee, J.W., 2005. Kinetics of absorption of Carbon Dioxide Wang, F., Wang, L.M., Wang, S.Q., Fu, D., 2017. Study on viscosity of MDEA-MeOH
in monoethanolamine solutions of polar organic solvents. J. Ind. Eng. Chem. 11, aqueous solutions. IOP Conference Series: Earth and Environmental Science 59,
202–209. 012020.
Park, S.-W., Lee, J.-W., Choi, B.-S., Lee, J.-W., 2006. Absorption of carbon dioxide into Xu, H.-J., Zhang, C.-F., Zheng, Z.-S., 2002. Solubility of hydrogen sulfide and carbon
non-aqueous solutions of N-methyldiethanolamine. Korean J. Chem. Eng. 23, dioxide in a solution of methyldiethanolamine mixed with ethylene glycol. Ind. Eng.
806–811. Chem. Res. 41, 6175–6180.
Patiño-Echeverri, D., Hoppock, D.C., 2012. Reducing the energy penalty costs of post- Zhang, L.-L., Wang, J.-X., Xiang, Y., Zeng, X.-F., Chen, J.-F., 2011. Absorption of Carbon
combustion CCS systems with amine-storage. Environ. Sci. Technol. 46, 1243–1252. Dioxide with ionic liquid in a rotating packed bed contactor: mass transfer study. Ind.
Price, J.P., 2014. Effectiveness of Financial Incentives for Carbon Capture and Storage. Eng. Chem. Res. 50, 6957–6964.
Bluewave Ressources, LLC, McLean, Virginia, USA. Zhao, B., Sun, Y., Yuan, Y., Gao, J., Wang, S., Zhuo, Y., Chen, C., 2011. Study on corrosion
Rezazadeh, F., Gale, W.F., Rochelle, G.T., Sachde, D., 2017. Effectiveness of absorber in CO2 chemical absorption process using amine solution. Energy Procedia 4, 93–100.
intercooling for CO2 absorption from natural gas fired flue gases using mono- Zheng, S., Tao, M., Liu, Q., Ning, L., He, Y., Shi, Y., 2014. Capturing CO2 into the pre-
ethanolamine solvent. Int. J. Greenhouse Gas Control 58, 246–255. cipitate of a phase-changing solvent after absorption. Environ. Sci. Technol. 48,
Riddick, J.A., Bunger, W.B., Sakano, T.K., 1985. Techniques of chemistry. Organic 8905–8910.

42

You might also like