You are on page 1of 9

AQUEOUS PIPERAZINE/POTASSIUM CARBONATE FOR ENHANCED CO2 CAPTURE

J. Tim Cullinane, Babatunde A. Oyenekan, Jennifer Lu, and Gary T. Rochelle* Department of Chemical Engineering, The University of Texas at Austin, Austin, TX 78712 ABSTRACT Experimental and modeling results are presented for aqueous piperazine (PZ)/potassium carbonate (K2CO3) for enhanced CO2 capture from flue gas. CO2 removal by aqueous absorption/stripping is a mature technology currently used in the manufacture of ammonia and syngas. The use of aqueous K2CO3 promoted by PZ should enhance rates of CO2 absorption and reduce the heat requirement for stripping. The thermodynamic and rate properties of this new solvent have been measured in a wetted wall column. Simple models have been developed to estimate absorber and stripper performance. The new solvent, 5m K+/2.5m PZ (m mol/kg water), provides CO2 solubility and capacity comparable to 7m (30 wt%) monoethanolamine (MEA). The heat of CO2 absorption is less than that in MEA solvents (22 kcal/mol) and decreases from 16 to 9 kcal/mol as temperature increases from 40 to 80oC and rich CO2 vapor pressure increases from 100 to 5000 Pa. CO2 absorbs in 5m K+/2.5m PZ 1 to 5 times faster than into 7m MEA. This faster rate can provide a rich solution from the absorber that has CO2 vapor pressure 1.4 to 2.6 times greater than with MEA. The piperazine solvent can also be used to provide 90% CO2 removal with a rich loading and packing height that would give only 75% removal with MEA. Because 5m K+/2.5m PZ provides a richer solution and greater optimum capacity, the reboiler duty with 5m K /2.5m PZ in a simple absorption/stripping process can be 25 to 46 % less that that with 7m MEA. Additional heat savings up to 6% may be achieved by operating the stripper at greater pressure.
+

INTRODUCTION Efficient methods for the capture and sequestration of carbon dioxide need to be developed to address global climate change. CO2 removal by aqueous absorption/stripping is a mature technology currently used in the manufacture of ammonia and syngas. The process incorporates a circulated solvent that chemically absorbs CO2 in an absorber and is regenerated in a stripper with the application of heat (low-pressure steam). Nearly half of the point-source CO2 emissions in the United States are from coal-fired power plants. High operating costs are associated with absorption/stripping in power production due to the high-volume removal required. Currently, commercial absorber/stripper processes utilize aqueous K2CO3, aqueous amine, or some blended solvent. Each possesses unique characteristics and brings specific advantages to the process. The development of a more efficient solvent, 5m K+/2.5m piperazine, for this application is the focus of this work. Amine solvents typically possess a high, but limited, rate of absorption. Faster solvents can minimize the height of packing required and associated capital costs and pressure drop. Alternatively, faster solvents may allow the column to operate at a closer approach to equilibrium, reducing steam requirements in the stripper. Bishnoi [1] showed that PZ increases the absorption rate of CO2 into aqueous methyldiethanolamine by a factor of 10. The addition of PZ to aqueous monoethanolamine increases absorption rates by as much as a factor of 2.5 [2]. The heat of CO2 absorption is an important thermodynamic characteristic and is directly related to the regeneration energy required. Unfortunately, the reaction of CO2 with amines has a relatively high heat of absorption (20-25 kcal/mol). A low heat of absorption, as found in aqueous K2CO3, would allow the design of more energy-efficient processes, but the rate of absorption in these solvents is slow and inefficient. Laddha and

Corresponding Author: tel. 512-471-7230, fax 512-475-7824, email: gtr@che.utexas.edu

Danckwerts [3] and Mahajani and Danckwerts [4] evaluated a number of amine additives to enhance absorption rates by a factor of 1.3 to 1.9 in aqueous K2CO3 [3, 4]. Exxon [5, 6, 7] found that diethanolamine or a specific hindered amine promoted absorption rates in aqueous K2CO3 by a factor of two to five. Rates in promoted K2CO3 compare favorably with other aqueous amines. A final distinguishing performance characteristic is CO2 absorption capacity, which is defined by VLE behavior. Increasing the capacity may allow operation in a more favorable equilibrium region. High capacities are generally achieved by increasing the solvent concentration. Blends of K2CO3 and PZ demonstrate promise as a more effective solvent [8]. A high concentration of PZ provides fast absorption rates. Likewise, a high total solvent concentration achieves CO2 capacities comparable to a 7m MEA (30 wt%). A 2:1 ratio of K+ to PZ approximates an optimum mixture as determined by both rate measurements and a calculated heat of absorption. Because of solid solubility limitations, the maximum practical solvent concentration is 5.0 mK+/2.5m PZ (m gmol/kg water). Although PZ costs $2.20/lb compared to MEA at $0.76/lb, PZ should be more resistant to thermal and oxidative degradation. This paper presents experimental and modeling results with this solvent. Bench-scale measurements provide thermodynamics and rate data. Modeling gives estimates of absorber and stripper performance. On-going testing in a pilot plant will demonstrate the process and validate the models. BENCH-SCALE EXPERIMENTAL METHODS Equilibrium partial pressure and rate of CO2 absorption were determined using a wetted-wall column over a broad range of temperatures (40 to 110oC) and solvent compositions (0 to 6.2m K+ and 0 to 3.6m PZ). A detailed description of the apparatus and methods of analysis can be found in Bishnoi [1] or Cullinane [8]. The solvent, contained in a 1400 cm3 stainless steel reservoir, was circulated through the apparatus at 2.5 to 3.5 cm3/s. The liquid formed a smooth film on the outer surface of a cylinder, contacting a counter-current gas stream. The apparatus provided a well-defined contact area of 38 cm2. The gas flowed through an annulus with a hydraulic diameter of 0.44 cm and a cross-sectional area of 1.30 cm2. Average values of the gas and liquid film mass transfer coefficients were 3 x 10-10 mol/Pa-cm2-s and 1 x 10-2 cm/s respectively. After contacting the gas stream, the solvent was collected and returned to the reservoir. CO2 loading was quantified by acid addition with CO2 stripping by N2 into an infrared analyzer. The solvent and apparatus is maintained at a constant temperature by a circulating bath of silicon oil. Nitrogen and carbon dioxide flow rates into the apparatus were set by mass flow controllers. Gas rates were 4 and 7 L/min to minimize gas phase resistance. The gases were mixed and saturated with water at the operating temperature of the column prior to entering the contactor. During contact with the solvent, CO2 was absorbed or desorbed at partial pressures from 70 to 40,000 Pa. The gas stream was then dried and sent to an Horiba PIR-2000 CO2 analyzer. CO2 partial pressures and mass transfer rates were estimated from the gas phase material balance. CO2 solubility was determined by varying CO2 partial pressure to get both absorption and desorption at a given solvent loading. Piperazine speciation in various mixtures was found using proton nuclear magnetic resonance (1H NMR) spectra from a Varian INOVA 500 NMR. The chemical shift of protons for piperazine, piperazine carbamate, and piperazine dicarbamate was previously determined by correlation to carbon-13 spectra. The 1H spectra are useful for quantitative interpretation of peak areas; peak areas of the spectra were integrated to determine the amount of a PZ species. Samples were prepared by replacing 20% of the water with deuterium oxide (D2O). Chemicals used include potassium carbonate (99.6 %) and potassium bicarbonate (99.9 %) purchased from Mallinckrod and anhydrous piperazine (>99 %) purchased from Sigma-Aldrich, Inc. D2O (99.9 %), for use in NMR experiments, was obtained from Cambridge Isotopes Laboratories. Desired CO2 loadings were achieved by varying the ratio of K2CO3 and KHCO3 with a constant K+ concentration.

THERMODYNAMIC MODELING AND RESULTS A rigorous thermodynamic model for K+/PZ was developed based on the electrolyte NRTL model [9]. The model includes OH-, CO2, CO3=, HCO3-, and the five piperazine species shown in Figure 1. Bench-scale results from NMR and wetted-wall column experiments were used to regress important model parameters. The model has been shown to predict the speciation and VLE behavior of the solvent effectively over a broad range of conditions [10]. A portion of the experimental data and model results are discussed along with bench-scale experiments below. The speciation of PZ in loaded solutions has been determined by 1H NMR for aqueous PZ and K+/PZ mixtures from 20 to 60oC. Comparisons at similar CO2 vapor pressures indicate a significant difference in speciation with the addition of K+. Figure 2 shows the electrolyte NRTL speciation of 2.5m PZ/ 5.0m K+. The high K+ provides CO32/HCO3- to buffer the solution at a high pH and delay the protonation of the amine until high CO2 loading. This lessens the decrease of free amine with loading. Conversely, the apparent stability of the carbamate and dicarbamate is increased, each potentially accounting for 45 50 % of the total PZ. Because PZCOO- maintains reactivity with CO2, the amount of reactive species is increased. Figure 2 shows that the total concentration of reactive species (PZ + PZCOO-) decreases with loading but still represents more than 40 % of the amine at rich conditions in the absorber (10,000 Pa). In contrast, the fraction of free MEA will be less than 10 % at rich conditions.
CO32- + CO2 + H2O

2 HCO31.0 100 Pa PZ 1000 Pa 10,000 Pa

O H N PZ O H N N O N H

0.9 0.8

+ CO2

H N

N
Fraction of Total PZ

PZCOOO

O O

0.7 0.6 0.5 0.4 0.3 0.2 0.1 PZCOO


-

PZ + PZCOO

+ CO2

N O

N O

PZ(COO-)2 H+PZCOO-

PZ(COO-)2 H N N H H H N
+

N H

PZH+ O H N N O H N
+

PZH+

O N

0.0 0.3

0.4

0.5

0.6
+

0.7

0.8

H O H+PZCOO-

Loading (mol CO2/(mol K + mol PZ)

Figure 1: Chemistry of K2CO3/PZ solvent.

Figure 2: PZ Speciation in 5.0 m K+/2.5 m PZ at 60oC, predicted by electrolyte NRTL model regressed to NMR data.

An additional benefit of increased carbamate stability is the potential reduction of apparent amine volatility. PZ solutions (2.5m) are more dilute than MEA (7m). Because only half of the amine exists as free PZ at lean conditions (P*CO2 = 100 Pa), a reduction in amine vapor pressure is expected at absorber conditions. This should reduce PZ make-up costs and circumvent the need for post-absorber gas treatment. Figure 3 gives CO2 equilibrium measurements from the wetted-wall column. The electrolyte NRTL predictions result from regression of a much larger data set with variable solvent concentration [10]. The model has been fitted with VLE data up to 80oC. The model tends to over-estimate the CO2 vapor pressure at high temperatures. At absorber conditions, the model fits the data to within 20 %. The CO2 capacity of each solvent was estimated by taking the difference between the equilibrium loading at 100 Pa of CO2 partial pressure and at 5000 Pa, typical of the gas concentration range of a coal-fired power plant. The capacity of 7 m (30 wt%) MEA at 60oC was estimated to be 1.40 mol CO2/kg H2O [11, 12]. Likewise, the capacity for 5.0 m K+/2.5 m PZ at 60oC was calculated to be 1.35 mol CO2/kg H2O, equivalent to MEA at these conditions.

10000

80

C
o

Table 1: Estimated heats of desorption (kcal/mol CO2) of 5.0 m K+/2.5 m PZ PCO2* at 40oC (Pa)
o

60

C C

40oC 17 16

80oC 9 12

40
1000

100 5000

PCO * (Pa)

100

10 0.4

0.5

0.6

Loading (mol CO2/(mol K + mol PZ))

0.7

0.8

Figure 3: CO2 Solubility in K+/PZ measured in the wetted wall column. Lines given by electrolyte NRTL model. Using the temperature dependence inferred from VLE predicted by the electrolyte NRTL model, a heat of CO2 desorption from K+/PZ has been calculated and is shown in Table 1. The values undergo a substantial change with temperature, varying by a factor of 2 from 40 to 80oC. The heat of desorption has a weak dependence on loading (CO2 vapor pressure) at both temperatures. The predicted values are 15 to 50% lower than in 7m MEA (20 - 22 kcal/mol CO2). Absorption/Desorption Rates CO2 absorption rates into K+/PZ have been measured in a wetted-wall column at 40 to 100oC. The rates are reported as normalized flux, the ratio of CO2 flux to the difference between CO2 partial pressure at the gas/liquid interface and the equilibrium CO2 partial pressure in the bulk solution (P*CO2). Figure 4 shows that absorption rates in 5m K+/2.5m PZ are 1 to 5 times faster than in 7m (30 wt%) MEA. At 40oC, the difference is most striking at high PCO2* where MEA rates drop quickly while rates in the K+/PZ solvent decline more slowly, giving rates up to two to three times faster. The fast rate observed in MEA at low PCO2* (~20 Pa) is likely the result of speciation. At this loading, MEA will contain a larger fraction of free amine due to the high stability of PZCOO- under similar conditions. At 60oC, K+/PZ maintains its advantage over MEA at low and intermediate loading, giving rates 1.5 to 2 times faster. At 10,000 Pa, however, the rates appear to be approximately equal. This may be the result of losing reactive PZ to PZ(COO-)2. Overall, K+/PZ out-performs MEA over the range of interest. It has also been shown to possess rates faster than other promoted K2CO3 solvents [8]. ABSORBER PERFORMANCE Model Freguia developed a model of absorber/stripper performance with 7m MEA in AspenPlus [12, 13]. Solution thermodynamics were represented by the electrolyte NRTL with regression of the data by Jou and Mather [11]. The absorber was simulated by RateFrac modified to correctly represent CO2 absorption with fast reaction with MEA in the mass transfer boundary layer. The stripper was represented by RateFrac with equilibrium chemical reactions in the boundary layer. Freguia provides calculated results for 90 % CO2 removal from atmospheric flue gas at 55oC containing 9.4 % H2O and 12.35 % CO2. The absorber used 15 m of CMR#2 random packing and the stripper used 10 m of CMR#2. In 30% MEA the optimum lean loading was 0.21 mol CO2/mol MEA, giving a rich loading of 0.46 and stripper heat duty of 206 kJ/mol CO2.

The Freguia model has been modified to emulate the performance of 5m K+/2.5m PZ. The carbamate equilibrium constant for MEA is given by:
MEACOO + H2O MEA + HCO 3

ln K

MEACOO-

= - 0.52 -

2545 T

(1)

To emulate the lower heat of absorption 5m K+/2.5m PZ the value of this constant has been changed to:
ln K
PZCOO-

= - 8.4

(2)

This has the effect of reducing the heat of absorption by about 5 kcal/gmol while keeping the solution thermodynamics otherwise identical at 50oC. Therefore the modified model will predict that 5m K+/2.5m PZ has approximately the same capacity and CO2 vapor pressure as 7m MEA. This modified thermodynamics model approximates the data in Figure 3. The rate model has been adjusted by increasing the rate constant of MEA with CO2 by a factor of 4. This approximation assumes that the diffusion of reactants and products does not become a limiting mechanism in the faster piperazine solvent. Because the Freguia model gives flux as the square root of the rate constant, this has the effect of increasing the normalized flux by a factor of 2, approximately consistent with the rate data presented in Figure 4, regardless of the mechanism. Predicted Absorber Performance Table 2 gives results calculated with the Freguia model for 7m MEA and the modified Freguia model for 5m K+/2.5m PZ. The piperazine solvent can provide richer solution with a constant amount of packing. With 15 meters of packing, 5m K+/2.5m PZ is able to give 90% removal with a higher effective loading in the rich solution. This advantage is best expressed by comparing the CO2 vapor pressure over the rich solution, which is 60% greater with 5m K+/2.5m PZ. With only 6.1 meters of packing (20 ft), 5m K+/2.5m PZ gives 75% removal with a rich CO2 vapor pressure which is 3.8 times greater than that of 7 m MEA. The piperazine solvent can also improve CO2 removal with the same amount of packing. With 6.1 meters of packing it provides 90% CO2 removal at 60% greater rich CO2 vapor pressure than MEA requires for only 75 % removal.

kg', normalized flux (gmol/Pa-cm -s)

7.0 m MEA, 60 C (Dang, 2001)

5.0 m K+/2.5 m PZ, 60oC

o 1e-10 7.0 m MEA, 40 C

(Dang, 2001)

5.0 m K /2.5 m PZ, 40 C

1e-11 10

100
2

1000

10000

PCO * (Pa)

Figure 4: CO2 absorption rate measured in the wetted wall column.

TABLE 2: ABSORBER PERFORMANCE WITH 7 M MEA, 5M K+/2.5M PZ, CALCULATED BY FREGUIA MODEL, INLET GAS WITH 12.35 % CO2, 9.4 % H2O, AT 55OC, LEAN LOADING TO GET P*CO2=40 PA AT 60OC CO2Removal (%) 90 90 75 75 90 CMR#2 Packing(m) Solvent 15 15 6.1 6.1 6.1 K+/ PZ MEA K+/ PZ MEA K+/ PZ Capacity(m) 2.22 1.47 2.09 1.25 1.81 Rich[CO2]T (m) 6.16 3.28 6.03 2.88 5.75 P*CO2,rich (kPa) 6.7 4.2 4.5 1.2 1.9

STRIPPER PERFORMANCE Spreadsheet Model A simple spreadsheet model has been developed to predict stripper performance with 5m K+/ 2.5m PZ and with 7 m (30 wt%) MEA. The model uses an equilibrium reboiler and 7 stages with a Murphree efficiency of 40 % for CO2, 100 % for water, and 100 % for heat transfer. The selected CO2 efficiency is somewhat arbitrary, but has little effect because most of the calculations approach an equilibrium pinch. The model represents equilibrium and enthalpies with approximate relationships. The CO2 vapor pressure (kPa) at stripper conditions was represented by a linear equation with three adjustable constants. The constants (Table 3) were determined by regressing the points from the rigorous model for 5m K+/2.5m PZ given in Figure 2 and data from MEA equilibrium flashes in AspenPlus with the rigorous model developed by Freguia [13] from data of Jou and Mather [11]. CO2 vapor pressure (kPa) at 40 and 60oC was obtained directly from the rigorous model for 5m K+/2.5m PZ and the AspenPlus model for MEA. TABLE 3: CO2 VAPOR PRES. (KPA) AT STRIPPER CONDITIONS AS FUNCTION OF [CO2]T (M) AND T (K).

ln P = a + b * [CO 2]T
5m K+/2.5 m PZ a b c 12.80 2.57 7788

c T

7 m MEA 20.58 2.04 8612

The reboiler was assumed to be in CO2 equilibrium. At each stage the CO2 concentration was calculated with the partial pressure definition of Murphree efficiency (Emv):

Emv =

Pn Pn 1 Pn * Pn 1

(2)

where Pn, Pn-1 are the partial pressures of CO2 on stages n and n-1, and

Pn* is the equilibrium partial pressure of CO2 leaving stage n. The partial pressure of water (kPa) was calculated with the approximate relationship:

PH O = exp ln 100 +
2

10000 1

1 1.987 T 373

(3)

With the linear equilibrium model, the heat of CO2 desorption is a function of CO2 loading and temperature. The average heat of desorption required between 100oC and 120oC was used for the energy balances in the stripper. The heat of CO2 desorption for 5m K+/2.5m PZ and 7 m MEA was 14.3 and 17.1 kcal/mol CO2, respectively. The heat of vaporization of water was approximated as 10 kcal/mol. The heat capacity of the gas was neglected. The heat capacity of the solvent was approximated as 1 kcal/kg H2O-oC. The column was solved from the top down. At each stage the material balances and energy balance were solved to obtain CO2 concentrations, gas rate, and liquid rate. The reboiler duty per mol of CO2 was calculated as the sum of the heat of desorption of the CO2, the heat of vaporization of water and the sensible heat required to bring the rich solution to the temperature of the stripper:
Q ( kcal / gmol CO 2 ) = nco 2 Hdes + nH O Hvap + Lcp (Tb Tt )
2

(5)

Predicted Stripper Performance The simple spreadsheet model was used to calculate the reboiler heat duty required with a given rich and lean loading. For most calculations, the pressure was 160 kPa and the rich solution was preheated to 10oC cooler that the lean solution from the stripper. For all cases in figures 5 and 6, the lean loading was adjusted to minimize the steam requirement (kcal/mol CO2) for a specified rich loading. The above procedure was run with rich loadings that correspond to equilibrium CO2 pressures of 1.25, 2.5, 5, and 10 kPa when the absorber operates at 40 and 60oC. Figure 5 gives the minimum reboiler duty and optimum capacity (mol CO2 stripped/kg/water) as a function of CO2 rich loading (mol all CO2 species/kg water) for the two solvents. The expected rich loading would be 3.4 mol/kg H2O for MEA and 5.3 for K+/PZ. The reboiler duty decreases a factor of two over the practical range of rich loading. Over the range of rich vapor pressures used to generate these results, the optimum capacity of 5m K+/2.5m PZ is 0.22 m greater than that of 7m MEA. An added advantage of this system is that 5m K+/ 2.5m PZ has a faster rate, therefore the absorber can be operated at a closer approach to saturation and a greater rich partial pressure than 7m MEA.
70 Reboiler duty (kcal/mol CO2) 2.5 Optimum capacity (gmol CO2/kg H2O)

60

Capacity

50
7m MEA

1.5

40

5m K+/2.5 m PZ

30

Duty

0.5

20 2.5 3 3.5 4 4.5 5 5.5 Rich [CO2]T (gmol CO2/kg H2O))

Figure 5: Optimum stripper performance, 160 kPa, 10oC approach.

Figure 6 shows the minimum reboiler duty as a function of the CO2 vapor pressure of the rich solution at 40oC and at 60oC. As with rich loading, the reboiler duty decreases a factor of two over the range of rich CO2 vapor pressure. At a given rich vapor pressure, 7m MEA requires 18 to 34% more heat than 5m K+/2.5m PZ. However, because of the difference in mass transfer rates, 7m MEA must operate at a rich CO2 vapor pressure of 5 kPa when 5m K+/2.5m PZ can be run at 10 kPa. With these conditions, when the absorber runs at 60o, 5m K+/2.5m PZ will require 35 % less heat than 7m MEA. If the rich vapor pressure achieved by 5m K+/2.5m PZ is always twice that of MEA, the range of energy savings is 25 to 46 %. Figure 7 shows that the reboiler duty for 5m K+/2.5m PZ is reduced 0 to 6 % by increasing the stripper pressure (and T) from 160 to 300 kPa, because the CO2 vapor pressure increases faster with temperature than the H2O vapor pressure. This effect is less pronounced with richer feed. Because MEA solutions are subject to degradation by polymerization at higher temperature, it is usually not feasible to operate above 120oC. However, piperazine is not an alkanolamine and is not subject to the same mechanisms of degradation, so it should be possible to use it in the stripper at greater pressure.

45

Reboiler duty (kcal/mol CO2)

Reboiler duty (kcal/gmol CO2)

60
o

Duty
40

60 C

2 35

40 40 C
o

300 kPa
30

160 kPa

1.5

7m MEA 5m K+/2.5 m PZ

1 25

0.5

20 0 2 4 6 Rich PCO2*(kPa) 8 10

20 4 4.5

[CO2]T (m)

5.5

Figure 6:

Optimum reboiler duty, 160 kPa,Figure 7: Stripper performance at elevated pressure, 5m K+/2.5m PZ, 10oC approach 10oC approach

FUTURE WORK Additional bench-scale work will investigate thermodynamics and rates at stripper conditions, oxidative degradation of piperazine, solid solubility limits, piperazine volatility, solvent reclaiming, and corrosion. Preliminary testing suggests that PZ will be more resistant to oxidative degradation that MEA. The absence of a alcohol group should make PZ less susceptible to high temperature carbamate polymerization. Vanadium has been used to inhibit corrosion of carbon steel in pilot plant testing of this solvent in progress in a 16.8-inch absorber/stripper. Sulfate accumulation may be controlled by precipitating potassium sulfate. ACKNOWLEDGEMENTS This work was supported in part by DOE Cooperative Agreement DE-FC26-02NT41440, by the Separations Research Program at the University of Texas, and by a number of industrial sponsors. However, any opinions, findings, conclusions, or recommendations expressed herein are those of the authors and do not necessarily reflect the views of the DOE or other sponsors.

Optimum Capacity (m)

Capacity

2.5

REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. Bishnoi, S. 2000. Carbon dioxide absorption and solution equilibrium in piperazine activated methyldiethanolamine. The University of Texas at Austin. Dang, H. 2001. CO2 absorption rate and solubility in monoethanolamine/piperazine/water. M.S. Thesis, Department of Chemical Engineering, The University of Texas at Austin. Laddha, S.S., and P.V. Danckwerts. 1982. The absorption of carbon dioxide by amine-potash solutions. Chem. Eng. Sci. Vol. 37, No. 5: 665-7. Mahajani, V. V., and P. V. Danckwerts. 1983. The stripping of carbon dioxide from amine-promoted potash solutions at 100 deg C. Chem. Eng. Sci. Vol. 38, No. 2: 321-7. Sartori, G., and D. W. Savage. 1983. Sterically hindered amines for carbon dioxide removal from gases. Ind. Eng. Chem. Fund. Vol. 22, No. 2: 239-49. Savage, D. W., G. Sartori, and G. Astarita. 1984. Amines as rate promoters for carbon dioxide hydrolysis. Faraday Discussions of the Chemical Society Vol. 77: 17-31. Tseng, P. C., W. S. Ho, and D. W. Savage. 1988. Carbon dioxide absorption into promoted carbonate solutions. AIChE J. Vol. 34, No.6: 922-31. Cullinane, J. T. 2002. Carbon dioxide absorption in aqueous mixtures of potassium carbonate and piperazine, M.S. Thesis, The University of Texas at Austin. Chen, C. C., H. I. Britt, J. F. Boston, and L. B. Evans. 1982. Local composition model for excess Gibbs energy of electrolyte systems. Part I: Single solvent, single completely dissociated electrolyte systems. AIChE J. Vol. 28, No. 4: 588-96. Cullinane, J. T., and G. T. Rochelle. 2004. The thermodynamics of aqueous potassium carbonate/piperazine for CO2 capture. Preprints of Symposia - American Chemical Society, Division of Fuel Chemistry Vol. 49, No. 1: 360-361. Jou, F.-Y., and A. E. Mather. 1995. The solubility of CO2 in a 30 mass percent monoethanolamine solution. Canadian Journal of Chemical Engineering Vol. 73, No. 1: 140-7. Freguia S. 2002. Modeling of CO2 removal from flue gases with monoethanolamine, M.S. Thesis, The University of Texas at Austin. Freguia, S., and G. T. Rochelle. 2003. Modeling of CO2 capture by aqueous monoethanolamine, AIChE J., Vol. 49, No. 7: 1676-1686.

You might also like