Conway 2014

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Article

pubs.acs.org/IECR

CO2 Absorption into Aqueous Solutions Containing


3‑Piperidinemethanol: CO2 Mass Transfer, Stopped-Flow Kinetics,
1 13
H/ C NMR, and Vapor−Liquid Equilibrium Investigations
William Conway,*,† Yaser Beyad,‡ Marcel Maeder,§ Robert Burns,§ Paul Feron,† and Graeme Puxty†

CSIRO Energy Flagship, Mayfield West, New South Wales 2304, Australia

CSIRO Energy Flagship, Clayton, Victoria 3169, Australia
§
Department of Chemistry, The University of Newcastle, Callaghan, New South Wales 2308, Australia
*
S Supporting Information

ABSTRACT: Global efforts to reduce carbon dioxide emissions stemming from the combustion of fossil fuels have
acknowledged and focused on the implementation of post combustion capture (PCC) technologies utilizing aqueous amine
solvents to fulfill this role. The cyclic diamine solvent piperazine has received significant attention for application as a CO2
capture solvent, predominantly for its rapid reactivity with CO2. A thorough investigation of alternative but simpler cyclic amines
incorporating a single amine group into the cyclic structure may reveal further insight into the superior kinetic performance of
piperazine and the wider applicability of such cyclic solvents for PCC processes. One such example is the cyclic monoamine 3-
piperidinemethanol (3-PM). To facilitate the evaluation of 3-PM as a capture solvent requires knowledge of the fundamental
chemical parameters describing the kinetic and equilibrium of the reactions occurring in solutions containing CO2 and 3-PM.
Additionally, in parallel with the preceding, experimental measurements of CO2 absorption into 3-PM solutions, including mass
transfer and vapor−liquid equilibrium measurements, can be used to validate the CO2 absorption performance in 3-PM solutions
and compared to that of monoethanolamine (MEA) under similar conditions. The present study is focused in two parts on (a)
determination of fundamental kinetic and equilibrium constants via the analysis of stopped-flow kinetic and quantitative
equilibrium measurements via 1H/13C nuclear magnetic resonance (NMR) spectroscopy and (b) experimental measurements of
CO2 absorption into 3-PM solutions via wetted wall column kinetic measurements, vapor−liquid equilibrium measurements, and
corresponding physical property data including densities and viscosities of the amine solutions over a range of concentrations and
CO2 loadings. Fundamental kinetic rate constants describing the reaction of CO2 with 3-PM are significantly faster than MEA at
similar temperatures (3-PM = 32 × 103 M−1 s−1, extrapolated to 40 °C from kinetic data between 15.0 and 35.0 °C; MEA = 13 ×
103 M−1 s−1, 40 °C). Conversely, the equilibrium constants describing the reaction between bicarbonate and amine, often termed
carbamate stability constants, are significantly lower for 3-PM than MEA at similar temperatures. Overall CO2 absorption rates in
3.0 M solutions of 3-PM and MEA, assessed in overall CO2 mass transfer coefficients, are lower in the former case over the entire
range of CO2 loadings from 0.0 to 0.4 mol of CO2 per mol of amine. The reduced absorption rates in the 3-PM solutions can be
attributed to higher solution viscosities and thus corresponding reductions in CO2 diffusion. CO2 absorption and cyclic capacities
in 3.0 M solutions of 3-PM and MEA were found to be significantly higher in the case of 3-PM. The larger CO2 capacities are
attributed to the lower stability 3-PM carbamate and the formation of larger amounts of bicarbonate compared to MEA. Overall,
the larger CO2 absorption capacity, cyclic capacity, and rapid kinetics with CO2 position 3-PM as an attractive CO2 capture
solvent.

1. INTRODUCTION Currently, the most viable and technologically capable


process which can effectively achieve the required and
The generation and release of carbon dioxide stemming from
immediate reductions in CO2 emissions from fossil fuel
the combustion of fossil fuels for the production of electricity,
power generation is post combustion CO2 capture (PCC)
and the potential impacts (of these emissions) on the global with subsequent storage of the captured CO2 gas in geological
climate system, are among the most significant environmental formations. Alternatively, the capture CO2 gas can be utilized
challenges of the modern era.1 It is generally acknowledged that for the purposes of enhanced oil recovery3 or as feedstock in
the polluting processes contributing to these underlying the manufacture of high value chemicals.4 An overview of the
emissions must be addressed in order to ensure future threats, current status of PCC technology indicates two common
impacts, and disruptions, to the environment and society, are roadblocks: large capital costs relating to the manufacture of the
minimized.2 In response to these issues, a conscious global
research effort is underway targeting methods which are Received: August 11, 2014
positioned to contribute to significant and immediate Revised: September 9, 2014
reductions in the evolution of parasitic greenhouse gas Accepted: September 22, 2014
emissions. Published: September 22, 2014

© 2014 American Chemical Society 16715 dx.doi.org/10.1021/ie503195x | Ind. Eng. Chem. Res. 2014, 53, 16715−16724
Industrial & Engineering Chemistry Research Article

absorption equipment and large and parasitic energy require- cCO2 is the dissolved concentration of CO2 in the liquid
ment for the PCC process which demands up to ∼30% of the (aqueous) phase.14,15
base-load output of a power station when PCC is deployed Once dissolved into solution, a series of parallel reactions
(devoted to the operation of the capture equipment based on a ensue between CO2(aq) and water (H2O), hydroxide (OH−),
benchmark technology employing 30% (w/w) monoethanol- and amine (R1R2NH where R1 and R2 are typically side groups
amine solvent). 5−7 Any industrial realization of PCC made up of carbon/hydrogen/oxygen/sulfur/nitrogen atoms in
technologies will require improvements in the practical design, the case of primary and secondary amines) to form carbonic
size, and operation of current-generation PCC processes. Of acid (H2CO3), bicarbonate (HCO3−), carbonate (CO32−), and
these, the simplest improvements to date have arisen from the carbamic acid/carbamate (R1R2NCO2H/R1R2NCO2−), respec-
investigation of the chemical solvent. Higher efficiency and tively. The kinetic reactions are described in eqs 2−4.
rapid reacting solvents are favored in view of replacing typical
k1
capture solvents such as monoethanolamine (MEA), methyl- CO2 (aq) + H 2O XooY H 2CO3
diethanolamine (MDEA), 1-amino-2-methyl-1-propanol k −1 (2)
(AMP), ammonia (NH3), and piperazine (PZ).8,9 The latter
k2
example involving PZ, a cyclic diamine, has been extensively CO2 (aq) + OH− XooY HCO3−
investigated for its rapid reactivity with CO2.10,11 While k −2 (3)
kinetically attractive, several limitations are apparent including
k7
low aqueous solubility, limited CO2 capacity at low PZ R1R 2NH + CO2 (aq) XooY R1R 2NCO2 H
concentrations, and issues relating to operation of the solvent k −7 (4)
at high concentrations due to precipitation. Despite the Equilibrium reactions describing the protonation(s) of
promising kinetic potential of PZ, the preceding issues indicate chemical species in solution including those of hydroxide
that alternatives are required.11,12 In view of the promising (OH−), carbonate (CO32−), bicarbonate (HCO3−), carbamate
kinetic performance of PZ a selection of alternatives can be (R1R2NCO2−), and amine (R1R2R3N) are described in eqs
derived from the similar subfamily of “cyclic monoamines” 5−9.
which incorporate aspects of the parent piperazine structure. In
a recent screening study13 in which a series of 76 amines were K3
CO32 − + H+ ↔ HCO3− (5)
examined for their CO2 absorption rates and absorption
capacities, the cyclic monoamine 3-piperidinemethanol (3-PM) K4
was identified as an interesting and promising CO2 capture HCO3− + H+ ↔ H 2CO3 (6)
solvent. 3-PM shares a cyclic structure similar to that of PZ K5
incorporating a single amine moiety into the ring structure; OH− + H+ ↔ H 2O (7)
however, unlike PZ, a methanol group is additionally attached
K6
to the ring in the case of 3-PM. In view of the large absorption R1R 2NH + H+ ↔ R1R 2NH 2+ (8)
capacity and CO2 absorption behavior, the preceding work has
been extended here to include a fundamental investigation of K8
R1R 2NCO2− + H+ ↔ R1R 2NCO2 H (9)
the chemical reactivity in 3-PM solutions and investigations of
CO2 absorption in 3-PM solutions via wetted wall covering a An alternative pathway leading to the formation of the
range of concentrations and CO2 loadings. carbamate exists for primary and secondary amines via a
1.1. Outline. CO2 absorption into a series of 3-PM reaction with HCO3− as described in
solutions from 0.1 to 3.0 M and CO2 loadings in a 3.0 M 3- k9
PM solution from 0.0 to 0.4 mol of CO2 per mol of amine have R1R 2NH + HCO−3 XooY R1R 2NCO2− ( +H 2O)
been investigated using a wetted wall column gas−liquid k −9 (10)
contactor at 40 °C. CO2 solubility in a 3.0 M 3-PM solution The kinetics of this reaction has also been observed by 1H
from 40 to 80 °C has been investigated using a vapor−liquid NMR spectroscopy.16,17 The significantly slower rate of this
equilibrium apparatus. The fundamental kinetic reactivity of 3- reaction in competition with reactions of CO2 with hydroxide
PM in the absence of physical mass transfer phenomena, and in and amine (eqs 3 and 4) was found to have a negligible impact
equilibrium solutions, has been investigated using rapid on the analysis of absorption flux data obtained from
stopped-flow kinetic and 1H/13C NMR spectroscopic equili- absorption measurements in a wetted wall contactor and
brium measurements, respectively. kinetic stopped-flow measurements.18 Thus, the reaction does
1.2. Kinetic/Equilibrium Reaction Set. The absorption of not appear in the kinetic model for the analysis of stopped-flow
CO2 in aqueous amine solutions incorporates consecutive and wetted wall measurements. However, the equilibrium
contributions from the physical dissolution of CO2 (gas to constant for the pathway is typically used to describe the
liquid phase) together with chemical reactions occurring in the stability of carbamates and appears exclusively in the model for
liquid phase which act to chemically consume dissolved the analysis of 1H NMR data here.
CO2(aq). The former process is described and quantified by 1.3. Chemical Constant Data and Activity Coefficient
the temperature-dependent Henry’s constant in Corrections. Values for the kinetic and equilibrium constants
pCO for the reactions described in eqs 2, 3, and 5−8 were taken
HCO2 = 2
= 2.82 × 106e−2044/ T from the literature.19−21 Kinetic constants for the formation of
cCO2 (1) the carbamic acid in eq 4, and the carbamate stability constant
in eq 10, were regressed from the stopped-flow kinetic and
where HCO2 represents the Henry’s constant for CO2 solubility NMR equilibrium measurements here, respectively. Values for
in pure water at infinite dilution (kPa M−1), p is the partial the protonation constant of the carbamic acid, log K8, eq 9,
pressure of CO2 in the gas (vapor) phase above the liquid, and were determined in the kinetic analysis using the equilibrium
16716 dx.doi.org/10.1021/ie503195x | Ind. Eng. Chem. Res. 2014, 53, 16715−16724
Industrial & Engineering Chemistry Research Article

constant data for eq 10, log K9, determined in the independent respectively. The chemical shift of the TSP standard was
1
H NMR analysis, and the equilibrium constants log K1, log K4, assumed constant and unaffected by temperature over the
and log K7 (calculated as K7 = k7/k−7) from eqs 2, 6, and 4, narrow range of temperatures investigated in the NMR
respectively, via microscopic reversibility as K8 = K7K4/K1K9. measurements here. All samples were tightly capped and sealed
Activity coefficient corrections were applied in the with para film tape so as to minimize any loss of CO2 gas from
calculations using the Debye−Hückel approximation and a the samples during equilibration and measurement. The
reduced version of the specific interaction theory (SIT) for concentrations were chosen here to ensure the solution
charged species. The former was used in the analysis of CO2(aq) concentration did not exceed the solubility at normal
fundamental kinetic and 1H NMR measurement data while the pressure. Solutions were thermostated in a temperature
latter was used in the analysis of wetted wall column and controlled water bath until equilibrium was completely reached
vapor−liquid equilibrium (VLE) measurements. All calculated (determined when no further change in the spectrum was
values correspond to standard state conditions at infinite observed). The set point temperature was maintained internally
dilution at zero ionic strength. The Debye−Hückel and SIT within the instrument (±1 °C) during the acquisition of
expressions are shown in eqs 11 and 12. spectra.
2.2.2. Spectral Processing and Peak Integration. 1H and
Az i 2 μ 13
C NMR spectra were processed individually using Bruker
log γi =
1+ μ (11) Topspin 3.0.1.b software. A baseline correction (polynomial)
was applied to all spectra prior to the integration of peaks.
Az i 2 μ Spectral peaks corresponding to amine and carbamate were
log γi = integrated in the topspin software, and the total concentrations
1 + 1.5ρ−1/2 μ (12) of 3-PM and its carbamate calculated from the total
where γi is the activity coefficient, A = (1.8248 × 10 )/ (eT)3/2
6 concentrations of amine ([3-PM] tot ) and bicarbonate
is the Debye−Hückel law slope, ρ is the density of water (kg ([HCO3−]tot) in the solution. The average integrals of 13C
dm−3), μ is the ionic strength (mol dm−3), and zi is the charge signals for the aliphatic carbons in 3-PM and its carbamate were
of species i. used in the equilibrium analysis of the low concentration
samples at 25.0 °C for comparison to the 1H data.
2. EXPERIMENTAL SECTION 2.2.3. NMR 3-PM/Na2CO3/HCl Titration Samples. A series
of equilibrium samples were prepared in 5 mm diameter NMR
2.1. Chemicals. High-purity carbon dioxide gas (CO2, BOC tubes containing the TSP/D2O capillary insert by dosing
Gases Australia, 99.99% purity), N2 (Coregas, 99.99%),
various amounts of a concentrated HCl solution (4.0 M) into
potassium bicarbonate (KHCO3−, BDH), sodium carbonate
1.5 mL of an equilibrated carbamate solution initially
(Na2CO3, BDH), potassium hydrogen phthalate (AJAX),
containing 0.5 M 3-PM and 1.0 M Na2CO3. The series of
sodium hydroxide solid (Merck), 3-piperidinemethanol (3-
additions were designed to cover the approximate range of pH
PM, Suzhou Rovathin Pharmatech), and hydrochloric acid
(AJAX) were all used as obtained. The concentrations of conditions over which carbamates are typically observed (pH
sodium hydroxide and 3-PM were standardized by potentio- 12−8). A duplicate series of samples containing 1.0 M 3-PM
metric titrations. A stock solution of sodium hydroxide was and 1.8 M Na2CO3 dosed with various volumes of HCl were
initially titrated with potassium hydrogen phthalate (KHP) prepared and analyzed in parallel.
standard, and this solution was used to standardize the 2.3. Stopped-Flow Kinetics. Fast reaction kinetics were
concentration of the hydrochloric acid stock solution. A performed on an Applied Photophysics DX-17 stopped-flow
known amount of the standardized hydrochloric acid solution spectrophotometer equipped with a J&M Tidas MCS 500-3
was added to the 3-PM solution and back-titrated to high pH diode array detector.22 The absorbance changes of acid−base
with standardized sodium hydroxide solution. All solutions indicators, corresponding to changes in the pH of the solutions
were prepared using ultrahigh purity milli-Q water which was upon reaction of solutions in the stopped flow, were observed
further boiled to remove dissolved CO2 gas. over the wavelength region from 400 to 700 nm. All solutions
2.2. 1H/ 13C NMR Spectroscopy. 2.2.1. Acquisition. were thermostated and maintained at the set point temperature
Quantitative proton (1H) and carbon (13C) NMR spectra ((15.0−35.0) ± 0.1 °C) by a circulating Julabo F20 water bath.
were acquired on a Bruker Avance Ascend 600 NMR The exact temperature of the solution was monitored by a
spectrometer operating at a frequency of 600.213 MHz. Proton thermocouple located inside the stopped-flow absorption cell
1
H spectra were obtained as the average of 16 scans with a pulse block.
delay time (D1) of 1 s. Inverse-gated decoupled 13C spectra 2.3.1. Reaction of CO2(aq) with 3-PM and Decomposition
were obtained at a pulse angle of 30° (zgig30 pulse program, of 3-PM Carbamate. First, the formation of 3-PM carbamate
Bruker) as the average of 128 scans with a pulse delay time was investigated by reacting equal volumes of a CO2(aq)
(D1) of 60 s, corresponding to a value ∼5 times that of the T1 solution (4.35 mM), with a series of 3-PM solutions ([3-PM]0
relaxation time of the slowest relaxing carbon (typically = 1.0−10.0 mM), in the presence of 0.05 mM alizarin red S and
carbamate/HCO 3 − ). Additional one-dimensional (1-D) 12.5 μM thymol blue indicators, in the stopped flow. Initially,
DEPT135 13C and two-dimensional (2-D) HSQC spectra CO2(aq) solutions were prepared by bubbling a mixed gas
were measured and utilized in the assignment of spectral peaks containing CO2 and N2 into a water solution located in a small
corresponding to 3-PM and its carbamate. These spectra are sample reservoir above the drive syringe in the stopped flow.
available in Figures S1 and S2 in the Supporting Information. The initial concentration of CO2(aq) in this solution was
TSP (3-(trimethylsilyl)propionic acid-d4, sodium salt) dissolved determined from the CO2 partial pressure in the gas stream
in D2O in a sealed glass capillary insert was present in the NMR (adjusted by gas flow controllers), and the corresponding
tube acting as the external reference and locking agents, Henry’s constant for CO2 solubility in water. Ideal behavior and
16717 dx.doi.org/10.1021/ie503195x | Ind. Eng. Chem. Res. 2014, 53, 16715−16724
Industrial & Engineering Chemistry Research Article

uniform mixing of the gases in the gas stream was assumed logarithmic mean of the inlet and outlet CO2 partial pressures.
here. The linear slope of the NCO2 against PCO2 plot is equal to the
Second, the decomposition of 3-PM carbamate at low pH overall mass transfer coefficient, KG. The interested reader is
conditions was investigated by reacting a preequilibrated directed to Darde et al.24 and Puxty et al.18 for a comprehensive
carbamate solution with hydrochloric acid solutions in the description of the relationships between CO2 diffusion and
stopped flow. Initially, a carbamate solution containing 0.025 M chemical reaction during absorption of CO2 into an amine
3-PM and 0.05 M HCO3− was prepared and equilibrated for 24 liquid and the mathematical principles of the absorption
hours at the desired temperature. The composition of the processes.
equilibrated carbamate solution was determined quantitatively 2.5. Density and Viscosity. Density and viscosities of the
using 1H NMR spectroscopy prior to the stopped-flow amine solutions were determined using a combined Anton Parr
measurements. This equilibrated carbamate solution was DMA-38 density meter (±0.001 g/mL) and AMVn viscometer,
further reacted in the stopped flow with a series of hydrochloric at each of the required temperatures. Density and viscosity
acid solutions to establish the decomposition reaction kinetics. measurements were repeated in triplicate, and the resulting
Absorbance changes of 0.05 mM alizarin red S and 0.025 mM value is the average of these repeats. Density and viscosity data
methyl orange indicators, and subsequently changes in the pH as a function of 3-PM and MEA concentrations from 0.5 to 6.0
of the solutions due to the decomposition of 3-PM carbamate, M, and in 3.0 M 3-PM and MEA solutions and CO2 loadings
carbonic acid, and bicarbonate, and protonation of amine, were from 0.0 to 0.4 mol CO2/mol amine, are presented in Figures
monitored. S3−S6 in the Supporting Information.
The preceding series of reactions were repeated over a range 2.6. Vapor−Liquid Equilibrium. CO2 solubilities in 3.0 M
of temperatures from 15.0 to 35.0 °C to establish the solutions of 3-PM and MEA were determined using a Parr VLE
temperature dependence of the reactions. Measurements were apparatus from 40 to 80 °C.25 The equilibrium distribution of
repeated a minimum of four times to accommodate a statistical CO2 between gas and liquid phases in each of the amine
analysis of the regressed kinetic and equilibrium parameters solutions was investigated by sequentially dosing CO2 gas from
which were compared to the standard deviations produced in a Swagelok CO2 cylinder, into a known volume of amine
the software. Analysis of the kinetic and equilibrium data was solution (100 mL), contained in a sealed and thermostated
performed using ReactLab Kinetic and Equilibrium software glass reaction vessel (Parr). The precise amount of CO2 dosed
packages (www.jplusconsulting.com) and in-house extensions into the vessel was determined by variation in mass of the CO2
of the software written in Matlab. cylinder before and after dosing, which was suspended from an
2.4. Wetted Wall Column. Absorption of CO2 into analytical balance (Mettler Toledo pb4002-s). Initially, and
individual aqueous solutions of 3-PM and MEA was performed prior to CO2 dosing, the vessel was repeatedly charged with
using a wetted wall column gas−liquid contactor at 40 °C. ultrapure N2 gas to ensure all gas impurities were removed from
Briefly, a stainless steel column with an effective height and the gas lines and gas headspace of the vessel. Following the N2
diameter of 8.21 cm and 1.27 cm, respectively, was used here.23 flush, stirring of the amine solution was initiated and
CO2 absorption flux, NCO2, into approximately 0.6 L of amine maintained at ∼500 rpm until steady state conditions were
solution was measured over a range of CO2 partial pressures reached as indicated by constant gas pressure reading
spanning 1.0−20.0 kPa. The total liquid flow in the apparatus (Swagelok S model transducer) and constant temperature
was maintained at 121.4 mL·min−1 (2.02 mL·s−1) by a (thermocouple). Total gas phase pressure and temperature in
Masterflex peristaltic pump. The liquid flow was monitored the reactor were monitored throughout the measurements
using a calibrated liquid flow meter. The total gas flow rate in following each dosing of CO2, until steady state/equilibrium
the system was maintained at 5.0 L min−1, and the desired was again reached. Once steady, the proceeding CO2 dosing
composition of the gas (i.e., CO2 partial pressure) established was initiated. From the total reactor pressure at each steady
by variation of Bronkhorst mass flow controllers, calibrated for state, and CO2 dosing, the total amount of CO2 remaining in
CO2 and N2 gases, respectively. The composition of the gas the gas phase, subtracting contributions to the total reactor
stream entering and exiting the wetted wall column, expressed pressure from water vapor and initial N2, and the subsequent
in vol % CO2, was continuously monitored using a Horiba VA- CO2 loading in the amine liquid were calculated according to
eqs 13−15
3000 IR gas analyzer. The former was measured at each CO2
partial pressure while bypassing the wetted wall column with (Pvessel(tot) − PH2O(vap) − PN2)Vgas
the gas stream passing directly to the Horiba gas analyzer. nCO2(gas) = Z
RT (13)
2.4.1. CO2 Absorption Flux (NCO2) and Overall Mass
Transfer Coefficients (KG). The amount of CO2 absorbing into nCO2(liquid) = nCO2 (dose) − nCO2(gas) (14)
the amine liquid was determined from the relative amount of
CO2 in the gas stream entering (bottom) and exiting (top) the loading CO × α = nCO2(liquid)/namine
wetted wall column. The CO2 absorption flux, expressed as 2 (15)
millimoles of CO2 absorbed per second, per unit area (mmol· where nCO2(gas) is the moles of CO2 gas, Z is the compressibility
s−1·m−2) of contact between liquid amine and gas, was
determined over a range of CO2 partial pressures in each of of CO2, Pvessel(tot) is the total pressure as measured inside the
the amine solutions. The preceding procedure was repeated for vessel at steady state, PH2O(vap) is the vapor pressure of water,
each of the amine solutions including those preloaded with PN2 is the partial pressure of nitrogen, Vgas is the volume of the
different amounts of CO2. gas phase (total vessel volume−amine liquid volume in m3), R
Overall mass transfer coefficients, denoted here as KG, were is the gas constant (8.314 J·(mol·K)−1), T is the temperature
determined from plots of the absorption flux, NCO2, against the (K), nCO2(dose) is the moles of CO2 dosed, nCO2(liquid) is the moles
applied driving force, PCO2. The latter was determined as the of CO2 dissolved into the amine solution, and namine is the
16718 dx.doi.org/10.1021/ie503195x | Ind. Eng. Chem. Res. 2014, 53, 16715−16724
Industrial & Engineering Chemistry Research Article

moles of amine in solution. Given the large molecular weight

k denotes rate constant; K denotes equilibrium constant and is calculated as Ki = ki/k−i. Numbers in parentheses represents standard deviation in the last digit. MEA data included for comparison. bk(T)
ΔS° (J mol−1 K)
and polarity of 3-PM, and the presence of the polar hydroxyl

−40(10)

81(12)
−43(7)

17(1)
group which can participate in hydrogen bonding interactions,
it was assumed that the amine vapor pressure is too low to

van’t Hoffd
provide an observable contribution to the total measured
pressure.

ΔH° (kJ mol−1)

6.3(2) × 10−3
3. RESULTS AND DISCUSSION

−23(3)
−23(2)
−19(3)
3.1. Kinetics. 3.1.1. Stopped-Flow Studies. The complete
series of kinetic measurements, including measurements of

17
carbamate formation and decomposition, were analyzed in a
single global fit at each temperature. The resulting analysis

ΔS⧧ (J mol−1 K)
produced kinetic and equilibrium constants for the reversible

74(8)

112(8)
−47(3)
−4(7)
formation of 3-PM carbamate, k7 and k−7, eq 4, and the
protonation constant of 3-PM carbamate, K8, eq 9, at each

Eyringc
temperature. Corresponding equilibrium constants for the
carbamate pathway in eq 4 were also calculated from the

ΔH⧧ (kJ mol−1)


ratio of the formation and decomposition rate constants as K7 =

73(3)

96(3)
38(1)
61(2)
k7/k−7. Subsequent analysis of the kinetic constants and their
temperature dependencies using the Arrhenius, Eyring, and

Table 1. Kinetic and Thermodynamic Data Describing the Reversible Formation of 3-PM Carbamate/Carbamic Acida
van’t Hoff relationships resulted in standard activation energies,
Ea, enthalpies, ΔH⧧/ΔH°, and entropies, ΔS⧧/ΔS°, of the
reactions, respectively. The calculated kinetic and thermody-

1.15 × 1019
17

5.8× 1010
1.0× 1013
1.17× 10
namic data are presented in Table 1.

A
Arrheniusb
From the data provided in Table 1 all kinetic values were
found to increase with temperature. A comparison of the
kinetic constants, k7, and the equilibrium constant, K7, for 3-PM
76(3)

98(3)
41(8)
63(2)
and MEA reveals consistently larger values for 3-PM at similar
Ea

temperatures. The underlying reason for this trend relates to


the relative basicities of the amines, 3-PM being ∼1.0 unit
4
1.98(5) × 10

= A exp(−Ea/RT). ck(T) = (k′T/h) exp((ΔH⧧ − TΔS⧧)/RT). dK(T) = exp(−(ΔH°/RT) + (ΔS°/R)).


2.4(3) × 102
7.6(4) × 103
1.9(3) × 102
larger than MEA at similar temperatures. We have previously
35.0 °C

investigated the intimate relationships26 between the kinetic

7.06(4)
81(17)

6.6(1)
41(6)
and equilibrium constants of carbamates and the protonation
constant of the amine. It was typically found that amines with
larger protonation constants resulted in large kinetic constants
3

4.9(2) × 103

1.1(1) × 102
8.2(5) × 10

in interactions with CO2. The resulting data for 3-PM here are
25.0 °C

consistent with this trend where the larger protonation constant


7.28(2)

6.4(1)
73(5)

90(8)

54(4)

of 3-PM correspondingly renders it as a faster reacting amine by


promoting stronger interactions with CO2. This is largely due
to the apparent size of the lone electron pair on the nitrogen
3

2.7(1) × 103

1.6(3) × 102
2.6(4) × 10

(amine group) which, in turn, increases the effective


15.0 °C

nucleophilic reactivity of the amine group toward CO2


16.8(2)

7.52(3)

6.3(5)
35(1)

76(5)

molecules. Conversely, the stability of 3-PM carbamate, as


outlined in the equilibrium constants log K9, is substantially
lower than the corresponding MEA carbamate. Molecular
log (K8/M−1)

log (K8/M−1)
s )

k7 (M−1 s−1)
−1

explanations for the reduced carbamate stability in the case of


constant

K7 (M−1)
K7 (M−1)
k−7 (s−1)

k−7 (s−1)

3-PM are complex and may reside from a combination of


−1
k7 (M

electronic and steric contributions, the latter forming the most


probable explanation where weak steric hindrance imposed by
the relatively close proximity of the large and bulky ethanol tail
3-PM

3-PM

3-PM
MEA

MEA

MEA

to the amine group, and thus the carbamate when formed, may
dissociate under particular conditions. Furthermore, it has been
suggested that the formation of intramolecularly hydrogen
RNH 2 + CO2 (aq) XoooY RNHCO2 H

RNHCO2− + H+ ↔ RNHCO2 H

bonded ring structures may contribute to decreased carbamate


stability.16 The Arrhenius activation energy for 3-PM
determined here is significantly larger than the corresponding
k−7

value for MEA and some other aliphatic amines, the values of
k7
reaction

K8

which are typically in the range of 40−65 kJ/mol. Chemical


explanations for the large values are unclear, however activation
energies reported in the literature for the cyclic amines such as
piperdine and pyrrolidine, 75.9 and 79.9 kJ/mol, respectively,
are similarly larger and close to the activation energies reported
here for 3-PM.27
a

16719 dx.doi.org/10.1021/ie503195x | Ind. Eng. Chem. Res. 2014, 53, 16715−16724


Industrial & Engineering Chemistry Research Article

3.1.2. CO2 Absorption Flux, NCO2, and Overall Mass attributable to the fact that the viscosity increases by a larger
Transfer Coefficient(s), KG. Overall mass transfer coefficients extent as the 3-PM concentration is increased, relative to MEA.
are critical parameters of amines in the assessment of their CO2 This increased viscosity inhibitively acts to lower the diffusion
absorption performance. Typically, KG values relate directly to coefficients of species in the liquid phase which, in turn, offsets
the absorption area and packing requirements for the capture the kinetic benefit of 3-PM and the larger reaction rate constant
process. Overall CO2 absorption rates into 3-PM and MEA in its reaction with CO2. Given the considerable structural
solutions over a range of amine concentrations and CO2 variations between 3-PM and MEA, specifically the large rigid
loadings from 0.0 to 0.4 mol of CO2 per mol of amine were cyclic ring structure in 3-PM, as opposed to the small free
investigated here at 40 °C using a wetted wall column. rotating chain structure of MEA, the observed increases in KG
Although CO2 absorption into MEA solutions has been are easily justified.
extensively investigated, a spread in published data relating to It should be emphasized that while it appears the amine
experimental variations justifies the repeated measurement of concentration has only a minor impact on KG at higher amine
the data here to allow a consistent comparison. Due to the concentrations (plateau region in Figure 1), considerable
limited solubility of 3-PM in aqueous solution the upper advantages in terms of the overall absorption uptake and
concentration was limited to 3.0 M. Similarly, measurements up CO2 capacity when operating the solvent at higher amine
to 3.0 M for MEA were completed and data from 3.0 to 5.0 M concentrations justifies the pursuit of more concentrated amine
were taken from our previous study and used in the solutions.
comparisons here. It should be noted that segments of 3.1.2.2. Effect of CO2 Loading on KG in 3.0 M Amine
nonlinearity were observed in the CO2 flux data for 3-PM. In Solutions. A comparison between the CO2 absorption
such cases, only the linear region of the flux curve was performance in 3-PM and MEA solutions, including solutions
considered in the calculation of KG values. over a range of CO2 loadings, was performed here. Selected KG
3.1.2.1. Effect of 3-PM and MEA Concentration(s) on KG data for individual 3.0 M solutions of 3-PM and MEA and for
Values. KG data for CO2 absorption into 3-PM and MEA CO2 loadings from 0.0 to 0.4 mol of CO2 per mol of amine are
solutions as a function of amine concentration, at zero CO2 presented in Figure 2. Overall, CO2 absorption is slower in the
loading, are presented in Figure 1. Despite larger kinetic

Figure 2. Overall mass transfer coefficients at 40 °C, KG, for 3.0 M


Figure 1. Overall mass transfer coeffeicients at 40 °C, KG, in 3-PM and solutions of 3.0 M 3-PM and MEA and for CO2 loadings from 0.0 to
MEA solutions as a function of amine concentration. 0.4 mol of CO2 per mol of amine.

constants for the reaction of CO2(aq) with 3-PM, the 3-PM solutions over the entire range of CO2 loadings.
corresponding trend from the stopped-flow data where 3-PM Although the KG values decrease with increasing CO2 loading,
is reacting significantly faster than MEA is reversed in the KG consistent with decreasing amounts of the free reactive amine
data. From Figure 1, CO2 absorption rates were found to be for reactions with CO2 in solution, the magnitude of the
comparable at low amine concentrations up to 1.0 M; however decrease is reasonably constant and linear with increasing CO2
the trend in KG increases more rapidly with increasing amine loading for both 3-PM and MEA. The observed decrease in KG
concentration in the MEA solutions beyond 1.0 M. with increasing CO2 loading is induced by the changing
Interestingly, a plateau in KG is observed along the sequences chemical and physical properties of the solutions with
at higher amine concentrations where KG increases only increasing CO2 loading; first the reduced availability of free
marginally with progressive increases of the amine concen- amine and second changes in the physical properties of the
tration above 2.0 and 3.0 M for 3-PM and MEA, respectively. solutions resulting from the increased concentration of ionic
Concomitant changes in the physical properties of the amine species (carbamate, bicarbonate, carbonate, and protonated
solutions as the amine concentration is increased are the amine) and the ensuing interactions of these species. From a
dominant causes of the effect. Any significant increases in KG theoretical perspective, and in order to compete with MEA at
with increasing amine concentration above the aforementioned similar concentrations and CO2 loadings, it is possible that the
concentrations are counterbalanced by parallel increases of the 3-PM solvent could be operated at slightly elevated temper-
solution viscosities which impact diffusion related processes atures which may result in a corresponding increase in the CO2
occurring in the solution. Further consideration should be given absorption rate. While the prospect of larger KG values at higher
to the preceding trends with respect to the individual amines. temperatures is attractive, the effect on diffusion and CO2
The slower overall CO2 mass transfer in the 3-PM solutions is solubility may, in parallel, impede such efforts. Similar
16720 dx.doi.org/10.1021/ie503195x | Ind. Eng. Chem. Res. 2014, 53, 16715−16724
Industrial & Engineering Chemistry Research Article

arguments can be made for the operation of MEA at elevated determination of fundamental chemical constants and their
temperatures; however the penalty to the absorption capacity, application to the prediction of realistic absorption flux data are
which is lower at higher temperatures, does not justify the broadly demonstrated.
pursuit for greater mass transfer rates. Conversely, given the 3.2. Equilibrium Studies. The equilibrium performance
larger absorption capacity of 3-PM, it can potentially sustain and CO2 capacity are critical properties in the selection of an
such a trade-off between larger KG values and CO2 capacity. amine solvent which ultimately defines the overall energy
3.1.3. CO2 Mass Transfer Predictions. Experimental demand of the desorption step. Two approaches are typically
absorption flux and mass transfer data have been predicted employed for the investigation of equilibrium behavior in amine
using the kinetic (rate) and equilibrium constants describing solutions; fundamental investigations of chemical equilibrium
the reaction of CO2 with 3-PM from the independent stopped- constants relating to the formation of carbamates and direct
flow kinetic study. A simple extrapolation of the kinetic measurements of CO2 solubility in a batch type reactor. The
constants to 40 °C via the Arrhenius expression was conducted former reveals intimate information about the energy related to
in the absence of measured values. An in-house software tool the formation (and destruction) of the carbamate (C−N
developed in Matlab was used for the prediction of molar CO2 bonds). Such information is additionally and fundamentally
absorption fluxes occurring in the amine film during absorption useful in the selection of novel solvents and in the prediction of
based on a comprehensive diffusion and chemical reaction equilibrium species distribution in CO2−amine solutions.
model.18,28,29 In the absence of measured values, diffusion Direct measurements of the equilibrium CO2 absorption
coefficients and CO2 solubility in the solutions were calculated capacity, as a function of increasing CO2 loading, based on
using correlations for these properties and the measured measurements of the equilibrium CO2 partial pressure in the
density and viscosity data.29 Experimental CO2 absorption data gas phase behavior, are principally simpler but do not carry
in the 3-PM solutions were predicted with the overall information about the chemical speciation in the solutions. We
agreement between the measured and predicted absorption have investigated both of these aspects here using 1H/13C
fluxes assessed in a parity plot as presented in Figure 3. NMR spectroscopy and direct measurements of CO2 solubility
via VLE measurements. The results of these studies are
presented and discussed in the following.
3.2.1. 1H/13C NMR Investigations: Carbamate Stability
Constants. The robust determination of equilibrium constants
describing the formation of carbamates, also commonly termed
carbamate stability constants, ideally requires a series of
measurements covering a range of pH conditions where the
carbamate is present in measurable and quantifiable amounts.
Simpler investigations which involve the extraction and analysis
of liquid samples from equilibrated amine solutions via NMR,
the solutions of which are typically prepared with increasing
amounts of CO2 sparged into the amine solutions, are prone to
significant error. In favor of this approach we have chosen to
perform a series of acid−base titrations of solutions containing
3-PM and Na2CO3, with various amounts of hydrochloric acid,
Figure 3. Parity plot of measured and calculated absorption flux in 3- to establish a series of samples covering the typical range of pH
PM solutions at 40 °C and zero CO2 loading. conditions where the carbamate is present and at a fixed amine
concentration and CO2 loading. This approach allows the
accurate preparation and delivery of both the amine and CO2
The agreement between measured and model calculated flux (as CO32− initially) to the sample without introducing small but
data is satisfactory in the amine solutions, in the absence of often significant errors associated with evaporative losses (both
preloaded CO2, over the entire range of 3-PM concentrations. solvent and water) during the CO2 bubbling and liquid
The average absolute deviation (AAD) over the entire series is sampling processes and gaseous loss of CO2 from the samples.
14.3%. Considering the dynamic nature of the experimental In an effort to establish the existence of any underlying
measurements where small variations in the expected CO2 effect(s) of concentration on the resolution and sensitivity of
loading of the solution from the mass difference method the NMR technique for the analysis of CO2/amine systems, we
translate into significantly large deviations in the modeling have prepared duplicate series of samples at both low and high
results, and the simple estimations for the film properties (film amine and CO32− concentrations, respectively. Additionally,
thickness, exposure time, and diffusion coefficients) in the selected samples have been analyzed in parallel via 13C
model, the result can be considered satisfactory. Overall, the measurements at 25.0 °C and the resulting integrals of the

Table 2. Equilibrium Constant Data for the Formation of 3-PM Carbamate Determined from the Analysis of 1H NMR
Measurements at 15.0−45.0 °C

van’t Hoff
ΔH° ΔS°
reaction constant NMR sample 15.0 °C 25.0 °C 35.0 °C 45.0 °C (kJ mol−1) (J mol−1 K)
K9 log (K9/M−1) 1
H high concn 0.93(1) 0.84(1) 0.76(1) 0.67(1) −14.8(1) −34(1)
RNH 2 + HCO3− ↔ RNHCO2− (+ H 2O)
1
H low concn 0.95(1) 0.86(1) 0.78(1) 0.71(1) −13.9(5) −30(2)

16721 dx.doi.org/10.1021/ie503195x | Ind. Eng. Chem. Res. 2014, 53, 16715−16724


Industrial & Engineering Chemistry Research Article

13
C spectra compared to the analysis of the corresponding 1H convolution of this behavior on the magnitude of stability
spectra. A typical series of 1H and 13C NMR spectra at 25.0 °C constants is highly complex and beyond the scope of this work.
in the equilibrium solutions containing 0.5 M 3-PM and 1.0 M The temperature dependence of log K9 here, expressed in the
Na2CO3 in the presence of additional amounts of HCl are ΔH° values, is within the range of values for similar cyclic
presented in Figures S7 and S8 in the Supporting Information. secondary amines we have assessed previously. The use of
Analysis of the concentration data obtained from the lower amine concentrations in our previous study, and the
integration of the carbamate and amine peaks in the NMR observation (or not) of carbamates in similar hindered amines
spectrum, together with the initial concentrations of the amine to 3-PM with alcohol groups in close proximity to the reactive
and Na2CO3 in the solutions, were analyzed using an amine group, further complicates any in-depth analysis of any
equilibrium chemical model incorporating reactions 5−10. trends.
The analysis resulted in values for the carbamate stability 3.2.2. Vapor−Liquid Equilibrium. Direct measurement of
constant, K9, at each of the temperatures from 15 to 45 °C. CO2 solubility in 3.0 M solutions of 3-PM and MEA at 40.0 to
Selected fits of the concentration data are shown in Figure S9 in 80.0 °C were performed using a vapor−liquid equilibrium
the Supporting Information. The resulting values for K9 at each apparatus. CO2 solubility data for a 3.0 M 3-PM solution is
of the temperatures, together with corresponding thermody- presented graphically in Figure 4.
namic values determined from subsequent analysis of the
temperature dependencies (of the equilibrium constants) via
the van’t Hoff relationship, are shown in Table 2.
The equilibrium constants and temperature dependences are
similar for the two series of solutions, confirming the effect of
concentration on the determination of equilibrium constants is
minimal here. Integrals determined from the analysis of 13C
data at 25.0 °C are similar to the corresponding 1H data and
were generally within ∼30% of each other. While the difference
between the 1H and 13C results may appear somewhat large,
the discrepancy relates to the sensitivity of the technique to the
13
C signal and relatively low abundance of the 13C isotope
(∼1% of the 12C isotope abundance). Thus, the integration of
small peaks close to the signal baseline is complex and prone to
slightly larger errors. The case is more relevant for amines such
as 3-PM where the carbamate is not present at significant Figure 4. Equilibrium CO2 partial pressure as a function of CO2
concentrations despite the use of high amine concentrations loading in a 3.0 M 3-PM solution from 40 to 80 °C. The solid black
here. However, both 1H and 13C NMR could be considered line corresponds to 15 kPa CO2 partial pressure (15% CO2).
suitable for the analysis of CO2/amine solutions where the
amount of carbamate formed is reasonably high so as to From the data in Figure 4 the gas phase CO2 partial pressure
minimize issues relating to the sensitivity of the NMR signal to increases marginally with increasing CO2 loading at low
carbamate concentration. A significant limitation of the NMR loadings and rapidly at high loadings above 0.3 at each of the
method here is the formation of CO2 gas bubbles resulting temperatures. The highest CO2 solubility is observed at low
from the reduced solubility at low pH. Thus, the investigation temperatures, which decreases significantly with increasing
of equilibrium behavior at significantly higher, and industrially temperature. Further changes in the CO2 partial pressure with
relevant, amine concentrations is not possible without increasing CO2 loading above ∼15 kPa are similar at each of the
significant modification of the procedure. While the conditions temperatures. While the quality and definition of the data at
here result in adequate and quantifiable amounts of carbamate, low CO2 loadings are somewhat discerning, related to the
the formation of considerable amounts of CO2 gas (bubbles) in limited sensitivity of the pressure measurements at very low
the solutions at lower pH conditions somewhat restrict the loadings below 20 kPa, the data presented here are useful for
method and analysis. Moreover, for this reason the inclusion of the purpose of defining the approximate absorption capacity of
a reaction defining the protonation of the carbamate into the the solvents. Such effects are commonly encountered in VLE
chemical model, K8, was found to have little impact on the measurements.
quality of the data fitting. Thus, this reaction is essentially not Absorption capacities, expressed on a moles per mole basis,
defined in the measurements here. at 40 °C and 15 kPa CO2 partial pressure (from the data),
3.2.1.1. Comparison of log K9 Values with MEA and Other together with estimated cyclic capacities calculated from the
Cyclic Amines. A comparison of the carbamate stability absolute difference in CO2 loading from 40 to 80 °C at 15 kPa
constants and thermodynamic values here with the correspond- CO2 partial pressure, are presented in Table 3. Absorption
ing values for MEA and a series of cyclic amines in our previous capacity at low temperature is some 20% larger for 3-PM than
work reveals several similarities.16 The average value (from the MEA. The larger capacity is a function of the low carbamate
low and high concentration NMR series) for log K9 = 0.84 at stability in the case of 3-PM which chemically favors the
25.0 °C here is considerably lower than the corresponding formation of larger amounts of bicarbonate and lower amounts
value for the unhindered parent cyclic amine piperidine (log K9 of carbamate at CO2 partial pressures similar to those of MEA.
= 1.38) and is aligned with the lower end of the range of values The result of this behavior is perpetually higher CO2 loadings
(log K9 = 0.79−2.69) for sterically unhindered cyclic secondary beyond 0.5:1.0 mol of CO2 per mol of amine can be achieved.
amines at similar temperatures. Considering the stability of 3- Moreover, the cyclic capacity of 3-PM over the temperature
PM carbamate is potentially influenced by the formation of an range here is almost double that of MEA for similar reasons. It
intramolecularly hydrogen bonded structure, any further is important to emphasize that any realistic comparison of the
16722 dx.doi.org/10.1021/ie503195x | Ind. Eng. Chem. Res. 2014, 53, 16715−16724
Industrial & Engineering Chemistry Research Article

Table 3. Estimated Absorption and Cyclic Capacities from NMR equilibrium plots, DEPT-135 and HSQC spectra, and
Experimental CO2 Solubility Measurements in 3-PM and calculated and measured equilibrium concentration profiles
MEA Solutions from 1H NMR analysis. This material is available free of charge
via the Internet at http://pubs.acs.org.


amine (3.0 M) absorption capacitya cyclic capacityb
3-PM 0.63/83c 0.21/28c AUTHOR INFORMATION
MEA 0.51/110c 0.11/24c
Corresponding Author
a
Moles of CO2 per mole of amine: 40 °C; 15 kPa CO2 partial
*E-mail: will.conway@csiro.au. Tel.: + 61249 606098.
pressure. bMoles of CO2 per mole of amine; difference in loading
between 40 and 80 °C at 15 kPa CO2 partial pressure. cGrams of CO2 Notes
per liter of solvent. The authors declare no competing financial interest.

maximum CO2 loading and cyclic capacities of amine solvents


should be performed on a mass basis (grams of CO2 per liter of
■ ACKNOWLEDGMENTS
This work was carried out within CSIRO’s Energy Flagship,
solvent), particularly in the case of solvents which will be and it was supported by the Australian Government through
operated at significantly different concentrations as is presented the Australia-China Joint Coordination Group on Clean Coal
here. The higher chemical efficiency of the 3.0 M 3-PM solvent Technology (JCG). The views expressed herein are not
is captured more effectively on a mass basis, and the overall necessarily the views of the Commonwealth, and the
capacity approaches that in a 5.0 M MEA solution. Commonwealth does not accept responsibility for any
Furthermore, on a mass basis, the larger cyclic capacity of 3- information or advice contained herein.
PM results in a larger overall CO2 cycle in 3.0 M 3-PM than in
5.0 M MEA. Thus, a process in which the two solvents would
be operated would require similar solvent inventories to achieve
■ REFERENCES
(1) Root, T. L.; Schneider, S. H. Conservation and Climate Change:
comparable cyclic capacities despite the significantly lower The Challanges Ahead. Conserv. Biol. 2006, 20, 706−708.
concentration in the former solution. This, in turn, results in (2) O’Neill, B. C.; oppenheimer, M. Dangerous Climate Change
advantages in terms of the capitol and operating costs related to Impacts and the Kyoto Protocol. Science 2002, 296, 1971−1972.
solvent makeup. Detailed process simulations will reveal the (3) Kim, Y. E.; Moon, S. J.; Yoon, Y. I.; Jeong, S. K.; Park, K. T.; Bae,
optimum operating temperature and CO2 loading ranges of the S. T.; Nam, S. C. Heat of Absorption and Absorption Capacity of CO2
solvent. in Aqueous Solutions of Amine Containing Multiple Amino Groups.
Sep. Purif. Technol. 2014, 122, 112−118.
4. CONCLUSION (4) Aresta, M.; Dibenedetto, A. Utilisation of CO2 as a Chemical
Feedstock: Opportunities and Challegnes. Dalton Trans. 2007, 2975−
The fundamental chemical and CO2 absorption behavior of 3- 2992.
piperidinemethanol has been investigated via a series of fast (5) Wang, M.; Lawal, A.; Stephenson, P.; Sidders, J.; Ramshaw, C.
kinetic stopped-flow, 1H/13C NMR equilibrium, wetted wall Post-Combustion CO2 Capture with Chemical Absorption: A State-of-
column, and vapor−liquid equilibrium measurements. The the-Art Review. Chem. Eng. Res. Des. 2011, 89, 1609−1624.
fundamental kinetic constants describing the formation of 3- (6) House, K. Z.; Harvey, C. F.; Aziz, M. J.; Schrag, D. P. The Energy
PM carbamic acid are significantly larger when compared to Penalty of Post-combustion CO2 Capture and Storage and Its
corresponding values for MEA at similar temperatures while Implications for Retrofitting the U.S. Installed Base. Energy Environ.
Sci. 2009, 2, 193−205.
carbamate stability constants are lower and thus highly favored. (7) Feron, P. Exploring the Potential for Improvement of the Energy
3-PM offers reasonable overall absorption rates compared to Performance of Coal Fired Power Plants with Post-Combustion
MEA at similar concentrations and loadings. The results here Capture of Carbon Dioxide. Int. J. Greenhouse Gas Control 2010, 4,
have additionally demonstrated that, in the absence of 152−160.
measured absorption flux data, the values can be reasonably (8) Aboudheir, A.; Tontiwachwuthikul, P.; Chakma, A.; Idem, R.
predicted using the fundamental chemical data including rate Kinetics of the Reactive Absorption of Carbon Dioxide in High CO2-
and equilibrium constants and corresponding physical property Loaded, Concentrated Aqueous Monoethanolamine Solutions. Chem.
data of the amine solutions. Significantly larger equilibrium Eng. Sci. 2003, 58, 5195−5210.
absorption and cyclic capacities were found for 3-PM over (9) Ramachandran, N.; Aboudheir, A.; Idem, R.; Tontiwachwuthikul,
MEA at similar conditions. P. Kinetics of Absporption of CO2 into Mixed Aqueous Loaded
Solutions of Monoethanolamine and Methyldiethanolamine. Ind. Eng.
From a strictly chemical point of view the rapid kinetics and Chem. Res. 2006, 45, 2608−2616.
low carbamate stability highlight 3-PM as an interesting (10) Bishnoi, S.; Rochelle, G. T. Absorption of Carbon Dioxide into
candidate for CO2 capture processes. The marginally slower Aqueous Piperazine: Reaction Kinetics, Mass Transfer, and Solubility.
absorption rates position 3-PM as a promising candidate for use Chem. Eng. Sci. 2000, 55, 5531−5543.
in a blended solvent which could potentially increase the (11) Freeman, S. A.; Dugas, R.; Van Wagener, D. H.; Nguyen, T.;
solubility of 3-PM and overall CO2 absorption rates. A more Rochelle, G. T. Carbon Dioxide Capture with Concentrated, Aqueous
comprehensive investigation of cyclic amines should be Piperazine. Int. J. Greenhouse Gas Control 2010, 4, 119−124.
undertaken to exhaustively determine if any further relation- (12) Rochelle, G. T.; Chen, E.; Freeman, S. A.; Van Wagener, D. H.;
ships (of these molecules) to the cyclic amine piperazine exist Xu, Q.; Voice, A. Aqueous Piperazine as the new Standard for CO2
and what is their overall suitability for PCC processes. Capture Technology. Chem. Eng. J. 2011, 171, 725−733.


(13) Puxty, G.; Rowland, R.; Allport, A.; Bown, M.; Burns, R.;
Maeder, M.; Attalla, M. Carbon Dioxide Postcombustion Capture: A
ASSOCIATED CONTENT Novel Screening Study of the Carbon Dioxide Performance of 76
*
S Supporting Information Amines. Environ. Sci. Technol. 2009, 43, 6427−6433.
Figures showing density and viscosity data in 3-PM and MEA (14) Versteeg, G. F.; Dijck, L. A. J. V.; Swaiij, W. P. M. V. On the
solutions and a range of CO2 loadings, stacked 1H and 13C Kinetics Between CO2 and Alkanolamines both in Aqueous and Non-

16723 dx.doi.org/10.1021/ie503195x | Ind. Eng. Chem. Res. 2014, 53, 16715−16724


Industrial & Engineering Chemistry Research Article

Aqueous Solutions. An Overview. Chem. Eng. Commun. 1996, 144,


113−158.
(15) Crovetto, R. Evaluation of Solubility Data of the System CO2−
H2O from 273 K to the Critical Point of Water. J. Phys. Chem. Ref.
Data 1991, 20, 575−589.
(16) Fernandes, D.; Conway, W.; Burns, R.; Lawrance, G.; Maeder,
M.; Puxty, G. Investigations of Primary and Secondary Amine
Carbamate Stability by 1H NMR Spectsoscopy for Post Combustion
Capture of Carbon Dioxide. J. Chem. Thermodyn. 2012, 54, 183−191.
(17) Conway, W.; Wang, X.; Fernandes, D.; Burns, R.; Lawrance, G.;
Puxty, G.; Maeder, M. Comprehensive Kinetic and Thermodynamic
Study of the Reactions of CO2(aq) and HCO3− with Monoeethanol-
amine (MEA) is Aqueous Solution. J. Phys. Chem. A 2011, 115,
14340−14349.
(18) Puxty, G.; Rowland, R.; Attalla, M. Describing CO2 Mass
Transfer in Amine/Ammonia MixturesNo Shuttle Mechanism
Required. Energy Procedia 2011, 4, 1369−1376.
(19) Wang, X.; Conway, W.; Burns, R.; McCann, N.; Maeder, M.
Comprehensive Study of the Hydration and Dehydration Reactions of
Carbon Dioxide in Aqueous Solution. J. Phys. Chem. A 2010, 114,
1734−1740.
(20) Harned, H.; Scholes, S. The Ionization Constant of HCO3−
from 0 to 50 °C. J. Am. Chem. Soc. 1941, 63, 1706−1709.
(21) Maeda, M.; Hisada, O.; Kinjo, Y.; Ito, K. Estimation of Salt and
Temperature Effects on Ion Product of Water in Aqueous Solution.
Bull. Chem. Soc. Jpn. 1987, 60, 3233−3239.
(22) Conway, W.; Wang, M.; Fernandes, D.; Burns, R.; Lawrance, G.;
Puxty, G.; Maeder, M. Toward the Understanding of Chemical
Absorption Processes for Post-Combustion Capture of Carbon
Dioxide: Electronic and Steric Consideration from the Kinetics of
Reactions of CO2(aq) with Sterically Hindered Amines. Environ. Sci.
Technol. 2013, 47, 1163−1169.
(23) Puxty, G.; Rowland, R.; Attalla, M. Comparison of the Rate of
CO2 Absorption into Aqueous Ammonia and Monoethanolamine.
Chem. Eng. Sci. 2010, 65, 915−922.
(24) Darde, V.; van Well, W. J. M.; Fosboel, P. L.; Stenby, E. H.;
Thomsen, K. Experimental Measurement and Modelling of the Rate of
Absorption of Carbon Dioxide by Aqueous Ammonia. Int. J.
Greenhouse Gas Control 2011, 5, 1149−1162.
(25) Richner, G.; Puxty, G. Assessing the Chemical Speciation during
CO2 Absorption by Aqueous Amines Using in Situ FTIR. Ind. Eng.
Chem. Res. 2012, 51, 14317−14324.
(26) Conway, W.; Wang, X.; Fernandes, D.; Burns, R.; Lawrance, G.;
Puxty, G.; Maeder, M. Toward the Understanding of Cheimical
Absorption Processes for Post-Combustion Capture of Carbon
Dioxide: Electronic and Steric Considerations from the Kinetics of
Reactions of CO2(aq) with Sterically Hindered Amines. Environ. Sci.
Technol. 2013, 47, 1163−1169.
(27) Garcia-Abuin, A.; Gomez-Diaz, D.; Navaza, M.; Vidal-Tato, I.
Kinetics of Carbon Dioxide Chemical Absorption into Cyclic Amines
Solutions. AIChE J. 2011, 57, 2244−2250.
(28) Puxty, G.; Rowland, R. Modeling CO2 Mass Transfer in Amine
Mixtures: PZ-AMP and PZ-MDEA. Environ. Sci. Technol. 2011, 45,
2398−2405.
(29) Wei, C.-C.; Puxty, G.; Feron, P. Amino Acid Salts for CO2
Capture at Flue Gas Temperatures. Chem. Eng. Sci. 2014, 107, 218−
226.

16724 dx.doi.org/10.1021/ie503195x | Ind. Eng. Chem. Res. 2014, 53, 16715−16724

You might also like