You are on page 1of 6

International Journal of Greenhouse Gas Control 12 (2013) 472–477

Contents lists available at SciVerse ScienceDirect

International Journal of Greenhouse Gas Control


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

Products and process variables in oxidation of monoethanolamine for CO2


capture
Alexander K. Voice, Gary T. Rochelle ∗
University of Texas, 1 University Station, C0400, Austin, TX 78712, USA

a r t i c l e i n f o a b s t r a c t

Article history: Oxidative degradation of monoethanolamine (MEA) in CO2 capture systems was studied using two dif-
Received 17 April 2012 ferent apparatuses and a variety of experimental conditions. Dissolved iron, manganese, and copper were
Received in revised form 2 November 2012 all potent catalysts, increasing the oxidation rate by up to a factor of ten. The activation energy of the
Accepted 13 November 2012
oxidation rate was 68–70 kJ/mol. Addition of 2% CO2 to the gas increased oxidation by a factor of five. The
Available online 9 January 2013
oxidation rate was found to be approximately first-order in oxygen partial pressure. Addition of 50 ppm
NO2 • or SO2 to the reactor gas had a minimal effect on the degradation rate. The mass balance was closed
Keywords:
within 5% at high gas flow and within 20% at low gas flow. Two-thirds of the degraded MEA in both
Monoethanolamine (MEA)
Oxidation
systems was converted to ammonia. Other major products were 1-(2-hydroxyethyl)-formamide (HEF)
Degradation and 1-(2-hydroxyethyl)-imidazole (HEI).
Ammonia production © 2012 Elsevier Ltd. All rights reserved.
Process variables

1. Introduction 2. Previous work

Oxidative degradation of amine solutions for CO2 capture is The presence of oxygen in gas-treating systems has long been
caused by the presence of excess oxygen for combustion in the recognized as a cause of degradation of MEA (and other amines),
flue gas entering the amine scrubber. Oxygen dissolves into the as well as production of corrosive degradation products (Hofmeyer
liquid phase and reacts irreversibly with the amine in the absorber et al., 1956; Lloyd and Taylor, 1954; Blanc et al., 1982; Dupart et al.,
packing, absorber sump, and cross-exchanger. Monoethanolamine 1993; Rooney et al., 1998; Blachly and Ravner, 1966, 1962). Previous
(MEA), a common amine for CO2 capture, was first observed by work has also established evidence of the chemical mechanism and
Kindrick to be particularly susceptible to oxidation (Kindrick et al., oxidative fragmentation products of MEA oxidation (Petrayev et al.,
1950). 1984; Denisov, 1996; Dennis et al., 1967; Russell, 1962; Rosenblatt
Solvent oxidation causes several problems for plant operators. et al., 1963) and how these mechanisms apply to CO2 capture
Degraded solutions have reduced CO2 capacity and may have (Rochelle and Chi, 2001). Later work focused specifically on rates
slower reaction rates compared with fresh solutions. In addition, of ammonia production and other oxidation products from MEA
heat stable salts resulting from oxidative degradation can increase solutions at absorber conditions (Chi, 2000; Chi and Rochelle, 2002;
corrosion. Make-up solvent must be added to replace degraded sol- Goff, 2005; Goff and Rochelle, 2004, 2006; Sexton, 2008; Sexton and
vent, and reclaiming processes must be used to remove degradation Rochelle, 2011).
products as they accumulate. Most importantly, many oxidation Blachly and Ravner (1966, 1962) first observed the catalytic
products belong to chemical classes known to cause environmental effect of dissolved metals present in MEA solutions on the oxida-
and health problems (Jackson and Attalla, 2010). One way to miti- tion rate. Recent work has further explored this effect by analyzing
gate these problems is to limit the amount of oxidative degradation for ammonia production as an indicator of MEA oxidation in a
occurring in the process and thus minimize the costs associated semi-batch laboratory system. Chi and Rochelle (Chi, 2000; Chi and
with degradation. Rochelle, 2002) demonstrated the role of iron as an oxidation cata-
lyst. Goff and Rochelle (Goff, 2005; Goff and Rochelle, 2004, 2006)
showed that copper alone or with iron was a more potent catalyst
than iron only.
∗ Corresponding author at: Department of Chemical Engineering, University of
Sexton and Rochelle (Sexton, 2008; Sexton and Rochelle, 2011)
Texas at Austin, 1 University Station, C0400, Austin, TX 78712, USA.
confirmed that copper was a potent catalyst by analyzing for
Tel.: +1 512 471 7230; fax: +1 512 471 7060.
E-mail addresses: voice@che.utexas.edu (A.K. Voice), gtr@che.utexas.edu
MEA loss and degradation product rates during prolonged solvent
(G.T. Rochelle). oxidation. Sexton also identified several new oxidation products

1750-5836/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijggc.2012.11.017
A.K. Voice, G.T. Rochelle / International Journal of Greenhouse Gas Control 12 (2013) 472–477 473

of MEA and discovered that ammonia, 1-(2-hydroxyethyl)-


imidazole (HEI) and 1-(2-hydroxyethyl)-formamide (HEF) were
the most prevalent products. Other likely products, which
have been reported in industrial systems and laboratory
experiments include 1-(2-hydroxyethyl)-oxalamide, N,N -bis(2-
hydroxyethyl)oxalamide, 4-(2-hydroxyethyl)-piperazin-2-one,
1-(2-hydroxyethyl) glycine and its corresponding amide N-(2-
hydroxyethyl)-2-(2-hydroxyethylamino)acetamide (Strazisar
et al., 2003; Lepaumier et al., 2009, 2011).
The effects of some process variables on MEA oxidation have also
been studied. Goff (2005) showed that the ammonia production
rate was higher in a lean (˛ = 0.15) solution than in a rich solution
(˛ = 0.40), and lowest in the absence of CO2 . Goff and Rochelle (Goff,
2005; Goff and Rochelle, 2004) also demonstrated that ammonia
production was first-order in oxygen concentration.
Supap et al. (2001) conducted oxidation of MEA at various Fig. 1. Anion chromatogram for a typical sample of oxidized MEA at absorber con-
concentrations in a batch reactor at high oxygen partial pressure ditions.
(241–345 kPa) and high temperature (120–170 ◦ C). The activation
energy of MEA oxidation was 66 kJ/mol, and the reaction was first- Program (SRP) Pilot Plant (PP) in Austin, TX (which was ana-
order in MEA and 1.5-order in oxygen. However, these results are lyzed and found to contain 0.5 mM Fe). Some experiments were
likely oxygen mass-transfer controlled, are confounded by changes also performed with 7 m MEA prepared from reagent-grade MEA
in oxygen pressure during the experiment, and contained unknown (Huntsman) with added dissolved metals. In addition, a high rate of
amounts of dissolved metal catalyst. gas sparged through the solution (HGF) or vigorous agitation (LGF)
This work has quantified the effects of process conditions, was used to increase oxygen mass transfer to the liquid.
as well as various gas- and liquid-phase contaminants, on MEA
oxidation—using a semi-batch system and representative condi-
tions. 4. Analytical methods

Samples taken from the liquid phase in the LGF and HGF reactors
3. Experimental methods were analyzed using a variety of methods, including anion chro-
matography, cation chromatography, HPLC–UV and total nitrogen.
This work used two semi-batch reactors designed to simulate Gas exiting the HGF was analyzed by hot-gas FTIR.
oxidation in the absorber packing in an MEA scrubbing system. The
low gas flow (LGF) system used 100 mL/min of 2% CO2 in oxygen 4.1. Anion chromatography
with reactor agitation at 1440 RPM. The high gas flow (HGF) used
2% CO2 in air sparged from the bottom of the reactor at a rate of Samples were prepared in two ways: dilution of the neat sample
7.65 L/min (dry basis). In the LGF, the water balance was controlled to 100× in water, and treatment of the sample with two volume-
by adding water daily to the marked level on the reactor; in the equivalents of 5 N sodium hydroxide, followed by 100× dilution
HGF the condenser and saturator were tuned to maintain the water in water after letting stand for 48 h at room temperature. NaOH
balance. treated samples allowed for recovery of organic acids (namely
The HGF experiment is similar to that described by Sexton formic acid and oxalic acid) from their respective amides via base-
(2008) and Goff (2005), with the exception that a higher reac- catalyzed hydrolysis (Figs. 1 and 2). This method was first used
tor temperature (70 ◦ C instead of 55 ◦ C) and oxygen concentration by Koike et al. (1987) for analysis of N-formyl-diethanolamine in
(21 kPa) was used to accelerate degradation. A condenser was also a 20% diethanolamine solution, and later by Sexton and Rochelle
used to reduce volatile amine loss from the reactor at 70 ◦ C. The (Sexton, 2008; Sexton and Rochelle, 2011) for analysis of various
low gas system used was the same as that described by Sexton and amide-degradation products in oxidized MEA solutions.
Rochelle (Sexton and Rochelle, 2011). The instrument was a Dionex ICS-3000 with IonPac AS15 ana-
The HGF was typically run for a period of one or two days to lytical column and mobile phase 2–45 mM aqueous potassium
achieve steady-state; NH3 production was used as an indication of
the relative amount of oxidation occurring. The LGF was run for
two weeks to achieve observable loss of MEA in the liquid-phase
and accumulation of liquid-phase degradation products.
It was initially assumed that ammonia accounted for a fixed per-
centage of degraded MEA; this percentage was later shown to be
roughly the same in the LGF and HGF. Since changes in the liquid-
phase MEA concentration require long experiment times, the major
advantage of using ammonia rates (in the HGF system) is that
changes in ammonia production rates can be observed instanta-
neously, thus significantly shortening the amount of time required
for an experiment.
Real process conditions were simulated by using low tempera-
tures (40–70 ◦ C) similar to those encountered in the absorber and
by using a continuous gas phase and batch liquid phase (to provide
a continuous supply of oxygen at roughly atmospheric pressure).
Most experiments were performed with 9 m MEA (PP), a sam- Fig. 2. Post-NaOH treatment anion chromatogram for a typical sample of MEA oxi-
ple (received in November 2010) from the Separations Research dized at absorber conditions.
474 A.K. Voice, G.T. Rochelle / International Journal of Greenhouse Gas Control 12 (2013) 472–477

Table 1
Analysis regions for FTIR.

Compound Range 1 (cm−1 ) Range 2 (cm−1 ) Range 3 (cm−1 )

Water 3157–3477 Fig. 3. Logistic equation.


CO2 910–1003 3425–3616 2165–2251
MEA 2416–3150
NH3 915–988 2423–2560
N2 O 2123–2224 2505–2628

Fig. 4. Derivative evaluated at t = 0.


hydroxide at 1.6 mL/min. Chromatograms for a typical sample are
shown below.
Table 2
Summary of NH3 rates from 7 m MEA (Huntsman) with metal additions. HGF con-
4.2. Cation chromatography ditions: 98 kPa air, 2 kPa CO2 , 70 ◦ C.

Metals added NH3 rate (mmol/kg/h)


Cation chromatography was used to determine liquid MEA, as
None (<1 ␮ Fe) 0.33
well as dissolved ammonia. Samples were diluted 10,000× in water 0.1 mM Fe 2.5
to reduce the MEA concentration to a manageable level. The instru- 0.1 mM Fe2+ + 0.01 mM Mn2+ 8.6
ment was a Dionex ICS-2100 with IonPac CS17 analytical column 0.1 mM Fe2+ + 0.25 mM Mn2+ 9.7
and aqueous methane sulfonic acid eluent (0–36 mM). This method 0.1 mM Fe2+ + 0. 50 mM Mn2+ 11.6
0.1 mM Fe2+ + 1.00 mM Mn2+ 12.6
allowed for fast separation of MEA from sodium and ammonium 1.0 mM Fe2+ + 1.00 mM Cu2+ 6.6
cations. 1.0 mM Fe2+ + 1.00 mM Mn2+ + 1.00 mM Cu2+ 11.6

4.3. HPLC–UV
5. Results
HPLC with a UV detector was used for determination of 1-(2-
hydroxyethyl)-imidazole in degraded MEA solutions. Samples were
5.1. Effect of metals
diluted 100× in water prior to injection on the HPLC. The instru-
ment was a Dionex U3000 with Dionex Polar Advantage 2 analytical
Aqueous metal sulfate salts were added to neat 7 m MEA in the
column and UV detector set to 210 nm. The mobile phase was ace-
HGF at 70 ◦ C with 98 kPa air and 2 kPa CO2 (Table 2 and Fig. 5). Rel-
tonitrile, methanol, and 10 mM aqueous ammonium carbonate. The
ative ammonia rates were used to determine the catalyst potency.
program started with 95% acetonitrile in 5% methanol and ended
Addition of 0.1 mM Fe2+ increased the ammonia production rate
with 50% acetonitrile in 50% aqueous ammonium carbonate. HEI
by more than a factor of six relative to the neat solution received
was eluted after approximately 13.5 min.
from Huntsman, Inc. (certified to contain less than 50 ppb of Fe) to
2.5 mmol/kg/h. Addition of 0.01 mM Mn2+ to this solution increased
4.4. Total nitrogen analysis the oxidation rate by 2.5× to 8.6 mmol/kg/h relative to iron only;
further additions of Mn2+ up to 1 mM increased the ammonia rate
Total nitrogen analysis was performed using an Aurora 1030 to 12.6 mmol/kg/h. Addition of 1 mM Cu2+ to a MEA solution at
Combustion TOC analyzer with total bound nitrogen (TNb ) module the same conditions resulted in an ammonia rate of 6.6 mmol/kg/h,
manufactured by OI Analytical. The sample was diluted to 3000× whereas 1 mM each of Fe2+ , Cu2+ , and Mn2+ resulted in an ammonia
in water; the injection volume was 0.2 mL. The sample is burned in rate of 11.6 mmol/kg/h.
a catalytic combustion tube before passing to a NOx converter that The results concerning the catalytic effects of iron and cop-
reduces NO2 to NO, which passes through an electrochemical NO per concur with previous work (Blachly and Ravner, 1966, 1962;
detector. Total nitrogen response decreases as MEA is oxidized to Chi, 2000; Chi and Rochelle, 2002; Goff, 2005; Goff and Rochelle,
ammonia, most of which exits the reactor with the gas. 2004, 2006; Sexton, 2008; Sexton and Rochelle, 2011). In this work,
manganese was found to be a catalyst of MEA oxidation of greater
potency than copper.
4.5. FTIR analysis

Gas exiting the HGF was analyzed using a Temet-Gasmet hot-gas


FTIR (DX-4000). Gas exiting the condenser on top of the HGF reac-
tor was pumped through a heated Teflon line maintained at 180 ◦ C 8
through a heated pump (also maintained at 180 ◦ C) to the anal-
NH3 Rate (mmol/kg/hr)

yser. The major components observed were water, CO2 , MEA, and 6
NH3 . A small amount of a fifth component, possibly N2 O, was also
observed in small quantities although this could not be confirmed.
No statistically significant amounts of NO, NO2 , formaldehyde, or 4
acetaldehyde were observed. A summary of the spectrum regions +0.01 mM Mn2+
used for each component is given in Table 1. 2

+0.1 mM Fe2+
4.6. Data analysis 0
0 5 10 15 20
Initial oxidation rates were determined by fitting the MEA con- Time (hrs)
centration as a function of time with an empirical fit (Fig. 3) and
Fig. 5. Oxidation of 7 m MEA with 98 kPa air and 2 kPa CO2 in the HGF at 70 ◦ C with
evaluating the derivative at time zero (Fig. 4). additions of Fe2+ and Mn2+ .
A.K. Voice, G.T. Rochelle / International Journal of Greenhouse Gas Control 12 (2013) 472–477 475

Fig. 8. Oxidation of 9 m MEA (PP) in the LGF with air and oxygen at 55 ◦ C and 70 ◦ C.
2% CO2 .
Fig. 6. Oxidation of 9 m MEA (PP) in the LGF with 98 kPa O2 , 2 kPa CO2 at 55 ◦ C.

5.2. SO2 and NO2 (after no more CO2 was being absorbed into the solution). Qual-
itatively, this result concurs with Goff’s observation (2005), that
The effect of NO2 and SO2 were studied by adding 50 ppm of ammonia production was lowest in the presence of no CO2 , and
each contaminant separately to the LGF with 98 kPa oxygen and higher at a lean-loaded condition that a rich-loaded condition. This
2 kPa CO2 at 55 ◦ C. The 9 m MEA (PP) solution containing 0.5 mM complex effect is likely due to two factors: at zero loading metals
Fe was used; MEA losses in the presence of SO2 and NO2 are com- are insoluble in the high pH MEA solution; at higher loading only
pared with that in the presence of oxygen and CO2 only (Fig. 6). the “free” MEA oxidizes (and not the MEA-carbamate or protonated
Neither SO2 nor NO2 had a significant effect on the loss of MEA. MEA). This explains the initial increase and subsequent decrease in
These three experiments demonstrate the reproducibility of the ammonia production with respect to CO2 loading. In an industrial
LGF experiment. The regressed initial rate of MEA loss in 9 m MEA process, the percentage of free MEA could range from 13% to 70%
from the SRP pilot plant is 29 ± 3 mmol/kg/h. This rate is signifi- (at 0.45 and 0.15 loading, respectively).
cantly higher than rates reported previously, and may be due to
the presence of certain dissolved metals in the solution (Kindrick
et al., 1950; Blachly and Ravner, 1966, 1962; Chi, 2000; Chi and 5.4. Oxygen dependence
Rochelle, 2002; Goff, 2005; Goff and Rochelle, 2004, 2006; Sexton,
2008; Sexton and Rochelle, 2011). MEA was oxidized in the LGF with air (98 kPa air dry = 17.9 kPa
O2 wet) and oxygen (98 kPa O2 dry = 83.3 kPa wet) at 55 ◦ C with 2%
5.3. CO2 dependence CO2 . MEA concentration is shown for both experiments in Fig. 8.
The initial rate of MEA loss (regressed) was 4.0 mmol/kg/h with air
The effect of CO2 on the ammonia evolution rate was studied and 25 mmol/kg/h with oxygen, indicating that the rate of MEA loss
by sparging unloaded, 7 m MEA with 2% CO2 in air in the pres- is approximately first order in oxygen. This generally concurs with
ence of 0.1 mM Fe at 70 ◦ C. Addition of CO2 to the reactor increased observations by Goff and Rochelle (Goff, 2005; Goff and Rochelle,
the NH3 rate from 0.54 mmol/kg/h to a maximum of 2.6 mmol/kg/h 2004) (who measured NH3 production rather than MEA loss) and
(at 0.30 loading) before falling to 2.3 mmol/kg/h at the final load- Supap et al. (2001) that MEA oxidation was around first-order in
ing (0.36) (Fig. 7). The loading was determined over the course of oxygen partial pressure. This also indicates that NH3 production is
the experiment by integrating the difference between the outlet a good indicator of relative degradation rates.
CO2 concentration and the equilibrium outlet CO2 concentration Goff and Rochelle (Goff, 2005; Goff and Rochelle, 2004) sug-
gested that MEA oxidation could be controlled by the absorption
rate of oxygen, based on weak effect of agitation with catalysis by
dissolved iron. However this observed activation energy is signif-
icantly greater than what might be expected for a mass transfer
controlled reaction.

5.5. Temperature dependence

The temperature dependence of MEA oxidation was determined


first by observing ammonia production rates at three tempera-
tures in the HGF and second by observing MEA loss over two
weeks in the LGF system at two temperatures and regressing the
initial rate (Fig. 8). In the first case, the activation energy was
70 kJ/mol, whereas in the second case it was 68 kJ/mol. All of the
rates observed in these experiments are shown in Fig. 9, where the
rates have been normalized by oxygen partial pressure.
The result is somewhat confounded by the fact that 2% CO2 was
Fig. 7. Oxidation of 7 m MEA with 98 kPa air and 2 kPa CO2 in the HGF with 0.1 mM
used in all experiments, thus the loading was lower at the higher
Fe at 70 ◦ C. temperature. Based on the above data on the effect of CO2 , a change
476 A.K. Voice, G.T. Rochelle / International Journal of Greenhouse Gas Control 12 (2013) 472–477

Fig. 9. Oxidation of 9 m MEA (PP) in the LGF and HGF at 55–70 ◦ C with 2% CO2 in
oxygen and air. Fig. 12. Products and material balance for oxidation of 9 m MEA (PP) in the LGF
(98 kPa O2 , 2 kPa CO2 , 70 ◦ C).

reactions between MEA and aldehydes. HEF is likely formed by


reaction of MEA with formaldehyde, followed by oxidation. HEI is
likely formed by reaction of ammonia, glyoxal, and formaldehyde
with MEA (Speranza and Wei-Yang, 1990). Significantly more HEI
was formed in the LGF reactor due to the fact that more of the
ammonia remains in the solution rather than being continuously
Fig. 10. Equations for the ammonia ratio in the LGF reactor (top), where TN(i) is the
(initial) total nitrogen concentration, and HGF (bottom). stripped out (as with the HGF). The relative amounts of some prod-
ucts may change in a real system that exposes the solvent to high
temperature in the stripper (Voice and Rochelle, 2012). However,
in loading from 0.40 to 0.35 at constant temperature could increase because ammonia always seems to represent more than 50% of the
the oxidation rate by 20%. nitrogen loss from MEA oxidation, it is a good surrogate to quantify
the rate of MEA oxidation.
5.6. Material balance
6. Conclusions
The material balance was closed to 100 ± 5% in the HGF and to
100 ± 20% in the LGF. Greater error is expected in the LGF because In conclusion, dissolved metals and temperature are the two
ammonia could not be analyzed directly. Control of the water bal- most important factors to consider in oxidation of MEA. Contami-
ance in the LGF reactor may have also contributed to greater error. nation by Fe and Mn from corrosion, fly ash, or other sources could
The ratio of ammonia products to lost MEA in the LGF and HGF cause up to a 38× increase in the oxidation rate of MEA in the
was calculated using the equations shown in Fig. 10. These equa- absorber, thus this is the most important factor to control. Based on
tions assume that disappearance of total nitrogen in the LGF is the regressed activation energy of 70 kJ/mol, the temperature bulge
primarily ammonia, that nitrate and nitrite are products of ammo- in the absorber (which can range from 40 to 80 ◦ C) could increase
nia oxidation, and that one ammonia equivalent is used to produce the oxidation rate by a factor of 19×. Decreasing the temperature
one equivalent of 1-(2-hydroxyethyl)-imidazole HEI (Speranza and bulge can provide significant benefit in reducing oxidation. Oxygen
Wei-Yang, 1990). concentration is set by the amount of excess air used for combus-
In the final sample, the percent of oxidized MEA recovered as tion to optimize power plant efficiency and therefore could not be
ammonia was 68% in the HGF and 71% LGF reactor (Figs. 10–12). changed for purposes of degradation. Low loadings may increase
For the LGF, it was assumed that the difference in total nitrogen oxidation by 6× compared with high loadings, however the load-
represented ammonia loss from the solution. Besides ammonia, ing is typically set to optimize energy performance of the capture
the two other major products are 1-(2-hydroxyethyl)-formamide plant. In addition, although higher loadings may benefit oxidation
(HEF) and HEI. Both of these products result from secondary by decreasing the free MEA, they would increase thermal degrada-
tion (Davis, 2009). The presence of SO2 and NO2 in the gas (up to
50 ppmv ) had no discernible effect on the oxidation rate of MEA.
The main oxidation products formed in the absorber (at low tem-
perature) are ammonia, HEI, and HEF with ammonia accounting
for about two-thirds of the degraded MEA. HEI production in an
industrial system is expected to be low (similar to the HGF system)
because ammonia is continuously stripped out.

Acknowledgement

This work was supported by the Luminant Carbon Management


Program.

References

Blachly, C.H., Ravner, H., 1962. The stabilization of monoethanolamine solutions for
Fig. 11. Products and material balance for oxidation of 9 m MEA (PP) in the HGF submarine carbon dioxide scrubbers. AD609888, NRL-6189. US Naval Research
(98 kPa air, 2 kPa CO2 , 70 ◦ C). Laboratory, Washington, DC.
A.K. Voice, G.T. Rochelle / International Journal of Greenhouse Gas Control 12 (2013) 472–477 477

Blachly, C.H., Ravner, H., 1966. Stabilization of monoethanolamine solution in carbon Lepaumier, H., Picq, D., Carrette, P., 2009. New amines for CO2 capture. II. Oxida-
dioxide scrubbers. Canadian Journal of Chemical Engineering 11, 401–403. tive degradation mechanisms. Industrial and Engineering Chemistry Research
Blanc, C., Grall, M., Demarais, G., 1982. The part played by degradation products in 48 (20), 9068–9075.
the corrosion of gas sweetening plants using DEA and MDEA. In: Laurance Reid Lepaumier, H., da Silva, E.F., Aslak, E., Grimstvedt, A., Knudsen, J.N., Zahlsen, K.,
Gas Conditioning Conference Proceedings, Norman, OK. Svendsen, H., 2011. 10th International Conference on Greenhouse Gas Control
Chi, Q.S., 2000. Oxidative Degradation of Monoethanolamine. M.S. Thesis. The Uni- Technologies. Energy Procedia 4, 1652–1729.
versity of Texas at Austin, Austin, TX. Lloyd, W.G., Taylor, F.C., 1954. Corrosion by and deterioration of glycol
Chi, S., Rochelle, G.T., 2002. Oxidative degradation of monoethanolamine. Industrial and glycol-amine solutions. Industrial and Engineering Chemistry 46 (11),
and Engineering Chemistry Research 41 (17), 4178–4186. 2407–2416.
Davis, J.D., 2009. Thermal Degradation of Aqueous Amines used for Carbon Dioxide Petrayev, E.P., Pavlov, A.V., Shadyro, O.T., 1984. Homolytic deamination of amino-
Capture. Ph.D. Dissertation. The University of Texas at Austin, Austin, TX. alcohols. Zhurnal Organicheskoi Khimii 20 (1), 29–34.
Denisov, E.T., 1996. Cyclic mechanisms of chain termination in the oxidation of Rochelle, G.T., Chi, S., 2001. Alkanolamine degradation. In: Rochelle, G.T., Chi, S.,
organic compounds. Russian Chemical Reviews 65 (6), 505–520. Bishnoi, S., Dang, H., Santos, J. (Eds.), Research Needs for CO2 Capture from Flue
Dennis, W.H., Hull, L.A., Rosenblatt, D.H., 1967. Oxidation of amines, IV. Oxidative Gas by Aqueous Absorption/Stripping. Final Report for U.S. DOE Contract DE-
fragmentation. Journal of Organic Chemistry 32, 3783–3837. AF26-99FT01029. Federal Energy Technology Center, U.S. Department of Energy,
Dupart, M.S., Bacon, T.R., Edwards, D.J., 1993. Understanding corrosion in alka- Pittsburgh, PA, pp. 103–135.
nolamine gas treating plants (Parts 1 & 2). Hydrocarbon Processing (May), 75–94. Rooney, P.C., Dupart, M.S., Bacon, T.R., 1998. Oxygen’s role in alkanolamine degra-
Goff, G.S., 2005. Oxidative Degradation of Aqueous Monoethanolamine in CO2 Cap- dation. Hydrocarbon Processing 77 (7), 109–112.
ture Processes: Iron and Copper Catalysis, Inhibition, and O2 Mass Transfer. Ph.D. Rosenblatt, D.H., Hayes, A.J., Harrison, B.L., Streaty, R.A., Moore, K.A., 1963. (I) The
Dissertation. The University of Texas at Austin, Austin, TX. reaction of chlorine dioxide with triethylamine in aqueous solution. Journal of
Goff, G.S., Rochelle, G.T., 2004. Monoethanolamine degradation: O2 mass transfer Organic Chemistry 28, 2790–2794.
effects under CO2 capture conditions. Industrial and Engineering Chemistry Russell, G.A., 1962. Peroxide pathways to autoxidation. In: Edwards, J.O. (Ed.), Per-
Research 43 (20), 6400–6408. oxide Reaction Mechanisms. Interscience Publishers, New York, pp. 107–128.
Goff, G.S., Rochelle, G.T., 2006. Oxidation inhibitors for copper and iron catalyzed Sexton, A.J., 2008. Amine Oxidation in CO2 Capture Processes. Ph.D. Dissertation. The
degradation of monoethanolamine in CO2 capture processes. Industrial and University of Texas at Austin, Austin, TX.
Engineering Chemistry Research 45 (8), 2513–2521. Sexton, A.J., Rochelle, G.T., 2011. Reaction products from the oxidative degrada-
Hofmeyer, B.G., Scholten, H.G., Lloyd, W.G., 1956. Contamination and corrosion in tion of monoethanolamine. Industrial and Engineering Chemistry Research 50,
monoethanolamine gas treating systems. In: National Meeting of the American 667–673.
Chemical Society, Dallas, TX. Speranza, G.P., Wei-Yang, S., 1990. Method for preparation of imidazoles. US Patent
Jackson, P., Attalla, M., 2010. Environmental impacts of post-combustion capture – 4,927,942.
new insights. 10th International Conference on Greenhouse Gas Control Tech- Strazisar, B.R., Anderson, R.R., White, C.M., 2003. Degradation pathways for
nologies. Energy Procedia 4, 2277–2284. monoethanolamine in a CO2 capture facility. Energy and Fuels 17 (4),
Kindrick, R.C., Atwood, K., Arnold, M.R., 1950. The relative resistance to oxidation of 1034–1039.
commercially available amines. Girdler Report No. T2.15-1-30. In: Report: Car- Supap, T., Idem, R., Veawab, A., et al., 2001. Kinetics of the oxidative degradation of
bon Dioxide Absorbents. Contract No. bs-50023, Gas Processes Division, Girdler aqueous monoethanolamine in a flue gas treating unit. Industrial and Engineer-
Corporation, Louisville, KY, 1950 (for the Navy Department, Bureau of Ships, ing Chemistry Research 40 (16), 3445–3450.
Washington, DC (Code 649P)). Voice, A.K., Rochelle, G.T., 2012. Sequential degradation of aqueous
Koike, L., Barone, J.S., Godinho, O.E.S., Aleixo, L.M., Reis, F.de A.M., 1987. N- monoethanolamine for CO2 capture. In: Recent Advances in Post-Combustion
formyldiethanolamine: a new artefact in diethanolamine solutions. Chemistry CO2 Capture Chemistry, January 1, 2012, pp. 249–263.
and Industry (September), 626–627.

You might also like