Vous êtes sur la page 1sur 9

Chemical Engineering Journal 210 (2012) 271279

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

A kinetic and process modeling study of CO2 capture with MEA-promoted potassium carbonate solutions
Hendy Thee, Yohanes A. Suryaputradinata, Kathryn A. Mumford, Kathryn H. Smith, Gabriel da Silva, Sandra E. Kentish, Geoffrey W. Stevens
Cooperative Research Centre for Greenhouse Gas Technologies (CO2CRC), Department of Chemical and Biomolecular Engineering, The University of Melbourne, Victoria 3010, Australia

h i g h l i g h t s
" MEA signicantly improves the overall absorption of CO2 into a K2CO3 solvent. " Reaction kinetics of CO2 absorption into a MEA-promoted K2CO3 solution is reported. " E-NRTL model suited for MEA-promoted K2CO3 system was developed using Aspen Plus. " MEA exhibits a larger rate constant at higher ionic strength conditions.

a r t i c l e

i n f o

a b s t r a c t
Aqueous solutions of carbonate salts such as potassium carbonate (K2CO3) have gained widespread acceptance as viable solvents for pre and post combustion capture of carbon dioxide (CO2). However, due to poor reaction kinetics a rate promoter is considered essential to improve the rate of CO2 absorption and hydration to bicarbonate. Using a well characterized wetted-wall column, we have evaluated the reaction kinetics of CO2 absorption into a K2CO3 solution promoted with monoethanolamine (MEA) under conditions resembling those found at industrial CO2 capture plants. Results presented here show that at 63 C the addition of MEA in small quantities, 1.1 M (5 wt.%) and 2.2 M (10 wt.%), accelerates the overall rate of absorption of CO2 in a 30 wt.% potassium carbonate solvent by a factor of 16 and 45 respectively. The Arrhenius expression for the reaction between CO2 and MEA is kMEA (M1 s1) = 4.24 109 exp(3825/T [K]) where the activation energy is 31.8 kJ mol1. Incorporating our experimental results into Aspen Plus, we have developed an E-NRTL model that can replicate pilot plant and simulate industrial capture processes employing K2CO3 promoted with MEA as the capture agent. 2012 Elsevier B.V. All rights reserved.

Article history: Received 20 June 2012 Received in revised form 15 August 2012 Accepted 16 August 2012 Available online 16 September 2012 Keywords: Potassium carbonate (K2CO3) Monoethanolamine (MEA) Carbon dioxide (CO2) Wetted-wall column Kinetics Process modeling

1. Introduction The removal and sequestration of carbon dioxide (CO2) from the ue gas of coal-red power stations by aqueous absorption is being actively investigated as a technology to help mitigate global warming due to human activities [13]. Aqueous solutions of both carbonate and amines, mainly monoethanolamine (MEA), are widely used in CO2 removal processes [2,46]. Patents from Germany exploring hot carbonate absorption of CO2 exist as early as 1904 [7]. In the 1950s Benson and Field developed the Beneld process which used hot potassium carbonate as a CO2 absorption solvent, with an objective of reducing the costs of synthesis-gas purication in which high CO2 partial pressures and high temperatures were employed [810]. This process was further developed during
Corresponding author. Tel.: +61 3 8344 6621; fax: +61 3 8344 8824.
E-mail address: gstevens@unimelb.edu.au (G.W. Stevens). 1385-8947/$ - see front matter 2012 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.cej.2012.08.092

the 1970s incorporating diethanolamine (DEA) as a rate promoter resulting in substantial lowering of capital and operating costs and higher treated gas purity [7]. In the carbonate system, CO2 is hydrated to bicarbonate HCO 3 via the following overall reaction (NB: in this work all species are aqueous unless otherwise stated) [11]:
CO2 CO2 3 H2 O ! 2HCO3

where the rate-limiting elementary reaction step is:

CO2 OH ! HCO 3

Under industrial CO2 capture conditions where pH is greater than 9, the contribution of the direct reaction between CO2 and H2O can be deemed negligible [12]. The rate of reaction rCO2 (d[CO2]/dt) is thus given by:

rCO2 kOH CO2 OH

272

H. Thee et al. / Chemical Engineering Journal 210 (2012) 271279

We have previously determined kOH using both a wetted-wall column (WWC) and stopped-ow spectrophotometry [11,13]. The vaporliquid equilibrium (VLE) to estimate the partial pressure of CO2 above the carbonate solution has also been previously studied and modeled [14]. In comparison to the carbonate system, amines have a relatively high rate of reaction with dissolved carbon dioxide. However, their performance as solvents is limited by a high heat of absorption, along with issues associated with amine loss and degradation and corrosion [15]. One way to improve the overall solvent system performance for CO2 capture is to blend a fast reacting amine, such as MEA, with a solvent that possesses a low heat of absorption, such as potassium carbonate (K2CO3), with the potential to take advantage of the benets of both solvents. The overall reactions between CO2 and a primary amine RNH2 can be represented as [16,17]:
CO2 RNH2 $ RNH 2 COO RNH 2 COO B $ RNHCOO BH

4 5

second step reaction in which the intermediate is deprotonated to produce the nal product (Eq. (5)). In the homogenous catalysis mechanism, both reactions take place at the gasliquid interface. In this study, we have characterized the reaction between CO2 and MEA within carbonate solutions and measured its activation energy under temperatures, pH levels and ionic strengths relevant to industrial carbon capture systems employing carbonate solutions. These results have implications for the operation of MEApromoted potassium carbonate solvent systems for carbon capture processes, and more specically allow more accurate design of absorber and regenerator units. Using this experimental kinetic data, a rate-based absorption model employing MEA/K2CO3 mixtures has been developed using Aspen Plus (version 7.3), which has been found to compare favorably with both bench and pilot scale data [5,25]. The Aspen Plus simulation also provides excellent predictive capability and is a very useful design tool for studying process variables such as temperature, pressure, CO2 loading and CO2 removal rate. 2. Materials and methods 2.1. Materials All chemicals employed in this study were of analytical reagent grade and used as supplied without further purication. Potassium carbonate (P99%, Thasco Chemical Co. Ltd.) and potassium bicarbonate (P99%, Sigma) were weighed to prepare a chemical equivalent of a 30 wt.% K2CO3 solution with known CO2 loading. MEA (P99%) was purchased from Science Supply Australia. CO2/N2 gas mixtures (10.2% and 89.8% CO2 in N2 obtained from BOC Gases Australia Limited) were used for all experiments and for the calibration of a MGA3000C CO2 gas analyser (ANRI Instrument and Control Pty. Ltd.). 2.2. Experimental methods The kinetics of CO2 absorption were studied using a wetted-wall column (WWC), a device that allows contact between a gas and a liquid phase with a controlled and measureable surface area for mass transfer, and thus, accurate measurements of the ux of CO2 into MEA-promoted potassium carbonate solvents. A detailed description of this apparatus and the characterization of its performance can be found in our prior work [11]. A process ow diagram of the full experimental setup is also provided in this reference. In summary, the WWC consists of a stainless steel tube with a total contact area of 4840 mm2. The chamber is housed inside a thick-walled steel chamber for temperature control. Liquid solvent ows up through the middle of the tube and is evenly distributed on the outer surface of the column. The liquid is collected at the bottom of the column and recirculated. The gas supply is pre-saturated with water and then ows counter-currently past the liquid lm before downstream concentration analysis. Liquid samples were analyzed by acid titration prior to and after experiments to determine solution pH and CO2 loading, which is dened as moles of CO2 absorbed per mole of K2CO3 [14]. The con2 centration of OH, HCO 3 and CO3 was determined using a Metrohm Titrando 809 autotitrator (Switzerland). During each experiment, data points were collected at a steadystate ux with a bulk CO2 partial pressure of 90 kPa and an initial solution loading of 0.15 (unless otherwise stated). Experiments were performed in which both MEA concentration and temperature were varied. Four repeat runs were carried out for each set of conditions and values reported here represent averages. Standard deviations from the repeat runs were used to calculate uncertainty intervals which are twice the standard error. As the rate of

The reaction shown in Eq. (4) results in the formation of a zwitterion intermediate and is rate limiting, while the reaction shown in Eq. (5) is the removal of a proton from the zwitterion by any base, B, to form the carbamate RNHCOO. At low CO2 loadings the species water, hydroxide ions and MEA itself can act as bases [18]. In pure MEA solutions, the reaction rate is thus given by:

r CO2 kMEA CO2 MEA

The CO2 absorption capacity of amine solvents is controlled by the consumption of the amine to form the carbamate. Furthermore, because of the high exothermicity of Eq. (5), the carbamate is also responsible for the high heat of absorption of amine solvents. These carbamates and their protonated counterparts can, however, undergo hydration reactions to regenerate the amine and produce a bicarbonate anion (Eq. (7)).

RNHCOO H2 O $ RNH2

HCO 3

The promoting effects of the addition of amines into carbonate solutions have previously been studied by several researchers [12,17,1921]. Usually primary or secondary amines are used as rate promoters while tertiary amines (where the zwiterrions can no longer deprotonate) do not show a signicant rate increasing effect [11,17]. These ndings are in agreement with the study of Laddha and Danckwerts [22] who observed an increase in the overall rate constants of reactions of CO2 with MEA and DEA upon the addition of small quantities of K2CO3. Versteeg et al. [17] summarized this literature and concluded that the addition of MEA to K2CO3 does result in an enhanced absorption rate. However, the kinetics of the reaction between CO2 and MEA in strong electrolyte solutions or highly CO2-loaded solutions which are normally encountered in industrial carbon capture systems and its process modeling were not well understood [2,6,23]. Researchers have considered two different chemical mechanisms to describe the catalytic effect of rate promoters in the carbonate system. A shuttle mechanism was proposed to describe the catalytic activity of primary and secondary amine promoters at low temperatures [12]. In this mechanism, the promoter acts as a carrier to provide another pathway for the transport of the absorbed CO2 from the gasliquid interface to the bulk liquid. However, a more recent publication by Savage et al. [20,24] indicated that the catalytic activity of amine in carbonate solutions can be better described by a homogeneous catalysis mechanism. The amine promoter in this mechanism acts as a homogenous catalyst for the reaction shown in Eq. (2) by rst forming an intermediate with the absorbed CO2 via Eq. (4). This process is followed by a very fast

H. Thee et al. / Chemical Engineering Journal 210 (2012) 271279

273

CO2 absorption was relatively small compared to the total CO2 concentration of the liquid, the equilibrium partial pressure of CO2 was assumed constant for each experiment, and was obtained from Endo [14]. The effect of the addition of MEA on the equilibrium partial pressure of CO2 was assumed to be negligible due to the high CO2 partial pressure used in the investigation. Experimental conditions have been chosen such that [OH] and [MEA] are in large excess over [CO2], resulting in pseudo-rst-order kinetics. In this case, the gas absorption rate can be expressed as [24]:

2.3. Modeling A simulation of a MEA-promoted potassium carbonate system was developed based on the rate-based absorption model of CO2 into MEA in Aspen Plus (Version 7.3). In this model, thermophysical property and reaction kinetic models were based on the work of Austgen et al. [35] and Hikita et al. [36]. We have incorporated a number of modications to the default model in order to achieve agreement with pilot plant results using a 32.5 wt.% aqueous MEA solution [25]. These modications include updating the rate of chemical reactions from experimental data and literature values, and adding binary interaction parameters between different components in the liquid phase. An open-loop absorption/desorption process owsheet diagram obtained from the default rate-based absorption model from Aspen Plus is shown in Fig. 1. The aqueous phase reactions occurring within the CO2MEA K2CO3H2O system can be classied into two categories: (1) those that are rate controlling and (2) those that are more rapid and thus can be assumed to rapidly reach equilibrium. The equilibrium equations are as follows:

NCO2

1 kg

1 p

P CO2 P CO2

2 ko l DCO2 kobs

NCO2 (mol m2 s1) is the ux of CO2 into MEA-promoted potassium carbonate solutions. PCO2 (Pa) is the bulk partial pressure of 3 CO2; P CO2 (Pa) is the equilibrium partial pressure of CO2. H (Pa m mol1) is the Henrys Law constant of CO2 in the solvent. kg (mol Pa1 m2 s1) represents the mass transfer coefcient in the  gas phase; kl (m s1) represents the physical liquid mass transfer coefcient. DCO2 (m2 s1) is the diffusivity constant of CO2 in the solvent. Here kobs (s1) is the observed pseudo-rst-order reaction rate constant, related to the CO2 reaction rate rCO2 (M s1) as given by:

RNHCOO H2 O $ RNH2 HCO 3


RNH 3 H2 O $ RNH2 H3 O

10 11 12 13

r CO2 kobs CO2

9a

Under our conditions the apparent rate constant kobs can then be expressed as:

2H2 O $ H3 O OH
2 HCO 3 H2 O $ H3 O CO3

K obs kOH OH kMEA MEA

9b

The Henrys Law constant and the diffusivity constant of CO2 in the potassium carbonate solutions were obtained from the literature [2631]. Data from this literature were used to conrm a correlation between the Henrys law constant in a potassium carbonate solution and ionic strength rst reported by Astarita et al. [32] and used by Proll et al. [31]. The same method was used when estimating the diffusivity constant of CO2 in the potassium carbonate solutions. This method of estimating the Henrys law constant and the diffusivity constant in a potassium carbonate solution based on its ionic strength was deemed acceptable considering that the amount of MEA in the potassium carbonate solutions was less than or equal to 10 wt.%. The physical liquid mass transfer coefcient was calculated using correlations developed by Pacheco [33] and Treybal [34]. The value of kOH was determined as a function of temperature using the Arrhenius expression kOH (M1 s1) = 2.53 1011 exp(4311/T [K]), obtained from our previous work in potassium carbonate solutions at similar ionic strength [11].

The equilibrium constants of these reactions are available in the literature and can be expressed as:

ln K eq a1 a2 =T a3 ln T a4 T

14

Coefcients a1a4 are summarized in Table 1 for these equilibrium reactions together with the applicable temperatures and the relevant literature references [35,37,38]. The rate controlling reactions are Eq. (2) and (4). The reaction rate constant (k) for these two equations is expressed in the form of an Arrhenius equation as follows:

K AeEa =RT

15

where R is the universal gas constant (kJ mol1 K1) and T is temperature (K). Pre-exponential factors (A) and activation energies (Ea, kJ mol1) for each reaction are summarized in Table 2. Rate constants for Eq. (2) were obtained from our prior work [11,13] in a temperature range of 4080 C and from the work of Pinsent et al.

GASOUT CO2OUT

LEANIN ABSORBER HEATER FLUEGAS TOHEATER RICHOUT PUMP LEANOUT


Fig. 1. Open-loop absorption/desorption process owsheet diagram within Aspen PlusTM.

RICHIN

STRIPPER

274

H. Thee et al. / Chemical Engineering Journal 210 (2012) 271279 Table 3 Speciation results for the wetted wall column experiment obtained from Aspen Plus. T (C) 43b [RNH2]added 0.5a 1.1 2.2 0.5 1.1 2.2 0.5 1.1 2.2 0.5 1.1 2.2 [RNH2]free 0.26 0.58 1.28 0.34 0.68 1.38 0.40 0.77 1.49 0.22 0.48 1.07 RNH 3 [RNHCOO] 0.03 0.06 0.10 0.04 0.06 0.11 0.04 0.07 0.12 0.04 0.08 0.16 HCO 3 h i CO2 3 2.73 2.85 2.95 2.62 2.71 2.80 2.50 2.56 2.62 2.28 2.45 2.63

Table 1 Temperature dependence of the equilibrium constants for reactions shown in Eqs. (7), (11)(13). Reaction 7 10 11 12 a1 3.4 140.9 220.1 6.7 a2 5851 13446 12432 3091 a3 0 22.5 35.5 0 a4 0 0 0 0 T (C) 40120 0225 0225 40120 Source [35,37] [38] [38] [35,37]

63b

Table 2 Pre-exponential factor (A) and activation energy (Ea) for reactions shown in Eqs. (2) and (4). Reaction 2 Direction A Forward Reverse 4 Forward Reverse
a a

83b

63c

Ea Reaction T (C) (kJ mol1) order 2 1 2 1 040 4080 040

Source [39] [11,13] [39]

0.28 0.41 0.54 0.21 0.31 0.43 0.14 0.22 0.32 0.32 0.49 0.69

0.60 0.46 0.32 0.72 0.59 0.46 0.84 0.75 0.63 1.38 1.17 0.91

4.32 1013 55.4 2.53 1011 35.8 2.38 1017 123.2 9.77 1010 2.62 108 3.23 1019 41.2 25.3 65.5

5.635.4 [36] 4383 This work 5.635.4 [36]

n.b. [K+] is not shown but allows charge balance. a The concentration unit of all species displayed in this table is k mol/m3. b The initial concentration of bicarbonate and carbonate ion before the addition of MEA is 0.85 M and 2.42 M respectively (CO2 loading = 0.15). c The initial concentration of bicarbonate and carbonate ion before the addition of MEA is 1.70 M and 1.99 M respectively (CO2 loading = 0.30).

The unit of the pre-exponential factor, A, varies depending on the order of the reaction. If the reaction is rst order, it has the unit s1. If the reaction is second order, it has the unit M1 s1.

[39] for 040 C. The kinetic parameters for Eq. (4) in a temperature range of 4383 C were taken from this work. The equilibrium constants used to calculate the kinetic parameters for the reverse of reactions shown in Eqs. (2) and (4) were obtained from the work done by Austgen et al. [35].

developed alternate binary interaction parameters for the K2CO3 KHCO3 system suitable for carbon capture processes [42]. In this investigation, a combination of the default parameters for MEA and the Cullinane parameters for K2CO3KHCO3 were used to simulate the CO2MEAK2CO3H2O system. 3. Results 3.1. Wetted-wall column kinetics

2.3.1. Binary interaction parameters The CO2MEAK2CO3H2O vaporliquid equilibrium was described using an Electrolyte Non Random Two Liquid (E-NRTL) activity model. Binary electrolyte pair parameters were used to predict the activity coefcient of each component in the system. Aspen Plus default parameters for MEA and H2O were used. These are obtained from the work of Austgen et al. [35] and have been shown to give an accurate prediction of VLE data for an MEA system [40,41]. The parameters are claimed to be applicable at a temperature up to 120 C and an MEA concentration up to 50 wt.%. However, such default parameters provide a more limited prediction for carbonate species [14]. Cullinane and Rochelle

12

43C 63C 83C

11

10

[MEA]Added / M
Fig. 2. Temperature-dependent effect of addition of MEA on the pH of a loaded 30 wt.% potassium carbonate solution.

Fig. 2 illustrates the effect of the addition of MEA on the pH of a 30 wt.% potassium carbonate solution based on experimental results. At 43 C the addition of 0.5 M, 1.1 M and 2.2 M MEA into the potassium carbonate solution results in an increase in the pH from 10.1 to 10.5, 11.0 and 11.8 respectively. These signicant increases in the solvent pH increase the rate of reaction for CO2 + OH and hence the overall absorption rate of CO2. Table 3 summarizes the speciation predicted by Aspen PlusTM under equivalent conditions including the concentration of free amine. Pseudo-rst order rate constants (kobs, s1) for CO2 reacting in MEA-promoted potassium carbonate at 43 C, 63 C and 83 C are shown in Fig. 3a. Results demonstrate that, at 63 C, the addition of MEA in relatively small amounts, 1.1 M (5 wt.%) and 2.2 M (10 wt.%), accelerates the pseudo-rst order rate of absorption of CO2 in a 30 wt.% potassium carbonate solvents by a factor of 16 and 45 respectively. It was also found that an increase in temperature improves the overall absorption of CO2. An increase in temperature from 43 C to 63 C and from 63 C to 83 C leads to an increase in the pseudo-rst order rate of absorption of CO2 in a 30 wt.% potassium carbonate added with 1.1 M MEA by a factor of 2.3 and 2.5 respectively. When the rate component due to CO2 + OH is excluded 0 kobs kMEA MEA kobs kOH OH , we observe non-zero rate constants, which can be attributed to the reaction of MEA with CO2. Fig. 3b shows a linear relationship between kobs and [MEA] indicating that the assumption of pseudo-rst-order kinetics is valid. The slopes of these lines provide values of the rate constant kMEA (M1 s1) for the CO2 + MEA reaction. This nding is in agreement with the work of Aboudheir et al. [43,44] who conducted their experiments over a range of MEA concentrations (0.19 3.88 M) and a lower temperature range (2040 C), and concluded that the order of reaction can be approximated to 1 with respect to the concentration of MEA. The same conclusion was also drawn by

pH

H. Thee et al. / Chemical Engineering Journal 210 (2012) 271279


150000

275

125000

43C 63C 83C

(a)

12

11

100000

ln [kMEA / M-1 s-1]

kobs / s-1

Versteeg et al. [17] This work Hikita et al. [36] Leder [45] Xiao et al. [47] Horng et al. [48]

75000

10

50000

25000

0 0.0 0.5 1.0 1.5


7 2.8 3.0 3.2 3.4 3.6

[MEA]free / M
150000

1000/T [K]
43C 63C 83C

(b)

125000

Fig. 4. Arrhenius plot of ln kMEA versus 1000/T for the reaction between MEA and CO2 from this work compared with the extrapolated Arrhenius t (dashed line) of a work by Versteeg et al. [17] and works by Hikita et al. [36], Leder [45], Xiao et al. [46], and Horng and Li [47].

100000

k'obs / s-1

75000

50000

25000

can be explained mainly due to the decrease in free amine concentration and the pH or hydroxyl ion concentration with increasing CO2 loading. Excluding the rate component attributed to the reac0 tion CO2 + OH, Fig. 6 shows the plot of kobs versus the concentration of free MEA. It can be deduced from this gure that loading 0 does not affect rate constant kobs when the concentration of active MEA is taken into account.
0.0 0.5 1.0 1.5

[MEA]free / M
Fig. 3. (A) Plot of pseudo-rst-order reaction rate constant kobs versus concentra0 tion of free amine [MEA]free at 4383 C. (B) Plot of kobs (=kMEA [MEA] = kobs kOH [OH]) versus [MEA]free at 4383 C. The presence of MEA improves the apparent 0 rate constant (kobs) especially at high temperatures. The reaction rate kobs increases linearly with [MEA] indicating a rst-order reaction rate. The slopes of these lines represent the rate constant kMEA.

3.2. Aspen plusTM model development and validation 3.2.1. CO2MEAH2O system The model was used to simulate a CO2 capture pilot plant that employed a 32.5 wt.% aqueous MEA solution as the capture agent [25]. Flue gas and inlet solvent compositions as well as the absorber specication in the simulation were matched to the pilot plant. The absorber had a total height of 11.1 m, an inside diameter of 42.7 cm and two 3.05 m packed beds with a collector plate. The packing selected for the simulation was IMTP No. 40, a random metal packing with a specic area calculated as 145 m2/m3. Mass transfer coefcients and interfacial area for IMTP No. 40 were predicted using the correlation presented by Onda [48]. Simulation results as well as experimental performance data over seven cases have been summarized in Table 4. Results show good agreement (deviation < 13%) between pilot plant data and simulation results. 3.2.2. CO2K2CO3H2O system The model was used to simulate data collected from a CO2 capture pilot plant operated at Hazelwood power station which employed a 30 wt.% un-promoted potassium carbonate solvent [6,49]. Inlet stream compositions and column specications were again matched to the pilot plant. The simulation results presented in Table 5 were found to be in agreement with the pilot plant data (deviation < 20%). The gas and solvent temperatures in the absorber and regenerator obtained from the simulation are higher than the pilot plant data as the simulation does not directly account for heat loss from the vessels. 3.2.3. CO2K2CO3H2O simulation compared to a large scale industrial CO2 capture plant The simulation model was also used to simulate an industrial scale CO2 capture process from a 600 MWe bituminous coal red

Versteeg et al. [17] who summarized works from other researchers and found the data to be consistent over a range of MEA concentrations (04.8 M) and a temperature range 040 C. For the MEA + CO2 reaction we determine the Arrhenius expression kMEA [M1 s1] = 4.24 109 exp(3825/T [K]) from 43 C to 83 C, where the activation energy is 31.8 kJ mol1. Fig. 4 shows that our measured rate constants are similar (deviation < 6%) to the extrapolated data from work done in a lower temperature range (5.6 C < T < 35.4 C) and MEA concentration (0.0152 M < [MEA] < 0.177 M) by Hikita et al. [36]. Good agreement (deviation < 5%) was also found between the experimental results obtained in this study and the model predictions provided by Versteeg et al. [17]. It is important to note that both work by Hikita et al. [36] and work by Versteeg et al. [17] were done in aqueous MEA without the presence of potassium carbonate. It was found that data from the present work are also in good agreement with the study conducted using a stirred vessel in the presence of potassium carbonate at a high temperature (i.e. 80 C) by Leder [45]. Data from the present work were also compared with those from more recent work done at lower temperatures by Xiao et al. [46] and Horng and Li [47]. It is important to note that the data from previously mentioned authors are in good agreement with the model predictions provided by Versteeg et al. [17]. Indeed the Versteeg correlation provides the best overall t to the data across the full temperature range. The pseudo-rst-order rate constant kobs was found to decrease with increasing CO2 loading in the liquid as shown in Fig. 5. This

276
120000

H. Thee et al. / Chemical Engineering Journal 210 (2012) 271279


120000

Carbonate Loading = 0.15 Carbonate Loading = 0.30


100000

(a)

Carbonate Loading = 0.15 Carbonate Loading = 0.30


100000

80000

80000

k'obs / s-1

kobs / s-1

60000

60000

40000

40000

20000

20000

0 0.0 0.5 1.0 1.5 2.0 2.5

0 0.0 0.5 1.0 1.5 2.0 2.5

[MEA]added / M
0

[MEA]free / M
Fig. 6. Plot of kobs versus concentration of free MEA at different carbonate loadings 0 at 63 C. Loading does not affect rate constant kobs when the concentration of active MEA is taken into account.

120000

Carbonate Loading = 0.15 Carbonate Loading = 0.30


100000

(b)

80000

60000

40000

20000

between MEA and ue gas impurities (such as SOx and NOx) which can form heat stable salts and subsequently reduce the capacity of the solvent to absorb CO2 are not included due to the scope of this work. MEA degradation due to industrial conditions such as high temperatures has also not been taken into account in the simulation model due to limited knowledge on the complicated degradation pathways.

kobs / s-1

0 0.0 0.5 1.0 1.5 2.0 2.5

4. Discussion
[MEA]free / M

Fig. 5. (a) Plot of kobs versus concentration of MEA added to a 30 wt.% potassium carbonate solution at different carbonate loadings at 63 C. (b) Plot of kobs versus concentration of free MEA at different carbonate loadings at 63 C. An increase in carbonate loading decreases the concentration of free amine in the solution and thus, the overall absorption of CO2 into MEA-promoted potassium carbonate.

power plant station [5]. The owrate and composition of the ue gas from the power plant are used as inputs to the simulation model. A 30 wt.% MEA solution is used to facilitate the absorption/ desorption process in the capture plant. The simulation results presented in Table 6 are found to match (deviation < 5%) the pilot plant data. 3.2.4. Model limitations It is essential to note that the simulation model has a number of limitations. The approach used only accounts for the major reactions involving the components MEA, K2CO3 and H2O. Reactions
Table 4 CO2MEAH2O simulation results compared to pilot plant data [25]. No. Lean loading [CO2]/[MEA] 1 2 3 4 5 6 7 8 9 0.28 0.29 0.23 0.23 0.23 0.28 0.28 0.28 0.28 Gas composition CO2 (mol%) 18.0 16.9 17.0 17.1 16.8 17.0 17.9 17.5 16.6 Gas rate

In this work, we have completed a detailed kinetic study of the CO2 + MEA reaction in potassium carbonate solutions under conditions found in industrial CO2 capture plants. Experimental results show that MEA signicantly catalyzes the overall absorption of CO2. A comparison of activation energy (Ea), pre-exponential factor 0 (A) and rate constant kobs between a number of amine-based promoters at 1.0 M and 40 C is presented in Table 7. The pre-exponential factor for this reaction, 4.24 109 M1 s1, is relatively small for a primary amine when compared to that of cyclic and secondary amine-based promoters, but is larger than a tertiary amine N-methyldiethanolamine (MDEA). The activation energy of 31.8 kJ mol1 is somewhat smaller than that of other amine-based promoters indicating that the reaction CO2 + MEA possesses a larger positive change in reaction rate at an elevated temperature. Additionally, we reveal that MEA exhibits a comparable rate constant (kobs , s1) to that of ethylenediamine and a larger rate constant than that of other amines from secondary and tertiary class (see Table 7), although it is approximately 10 times smaller than that of piperazine at 40 C.

Liquid rate (L/min) 30.1 60.8 39.4 56.8 83.1 42.8 42.6 40.7 54.9

CO2 removal (%) Experimental 69 86 72 87 94 95 80 95 70 Simulation 72 79 74 80 82 89 89 89 73

(m3/min) 8.2 8.2 11.0 10.9 11.0 5.6 5.5 5.5 11.0

H. Thee et al. / Chemical Engineering Journal 210 (2012) 271279 Table 5 CO2K2CO3H2O simulation results compared to the pilot plant at Hazelwood power station [6]. Unit Absorber Stream Gas-inlet Parameter Temperature, C Pressure, kPa g Flowrate, kg h1 Temperature, C Pressure, kPa g Flowrate, kg h1 Temperature, C Flowrate, kg h1 Temperature, C Flowrate, kg h1 Temperature, C Temperature, C Temperature, C Temperature, C Pressure, kPa g Flowrate, kg h1 Pilot plant 44.7 3.2 5602 38.1 0.3 4360 40.1 39870 43.6 39940 115 113 110.3 48.1 2000 Simulation Input Inputa Inputa 41.3 0 4383 Inputa Inputa 54.2 41089 Inputa Inputa 105.9 113.3 49.1 Inputa
a

277

Deviation (%) 8.4 0.5 19.6 2.9 2.7 2.1

Gas-outlet

Solvent-inlet Solvent-outlet Regenerator Column Solvent-inlet Solvent-outlet Vapor-outlet

Some of the pilot plant data were used as input data in the simulation model.

Table 6 CO2K2CO3H2O simulation results compared to a large scale CO2 capture plant using ue gas from a 600 MWe bituminous coal red power plant station [5]. Unit Absorber Stream Gas-inlet Parameter Temperature, C Pressure, kPa Flowrate, kg h1 CO2 removal,% Temperature, C Pressure, kPa Flowrate, kg h1 CO2 loading Temperature, C Pressure, kPa Flowrate, kg h1 CO2 loading Temperature, C Pressure, kPa Thermal energy, GJ/ton CO2 Plant data 50 101.6 2.18 106 90 50 180 7.22 106 0.242 51 101.3 7.44 106 0.484 108 130.0 4.24 Simulation Inputa Inputa Inputa 90 Inputa Inputa Inputa Inputa 49.5 101.6 7.47 106 0.479 103 131.3 4.46 Deviation (%) 0 +3 0.3 0.4 +1 +5 +1 +5

Gas-outlet Solvent-inlet

Solvent-outlet

Regenerator

Column

Some of the pilot plant data were used as input data in the simulation model.

Table 7 0 Activation energy, pre-exponential factor and rate constant kobs (s1) comparison of 1.0 M MEA and other amine-based promoters at 40 C. Class Cyclic Primary Promoter Piperazine Ethylenediamine MEA kobs a (s1) 10.3 104 35.6 103 13.1 103 11.7 103 10.5 103 21.0 103 3.5 103 1.4 102 2.8 10 1.3 10 1.2 103 p Ka b 9.7 10.0 9.5 Orderc 1st/2nd 1st/2nd 1st/2nd Ad 4.1 1010 3.1 1013 4.4 1011 9.8 1010 1.2 1011 8.5 1011 8.26 1015 3.0 1011 4.2 109 2.6 1012 5.2 1010 3.1 108 4.0 108 1.1 1010
0

Ea (kJ mol1) 33.6 53.6 44.9 41.2 42.2 46.7 71.0 44.7 31.8 53.1 51.5 42.4 44.9 41.7

Temp range (C) 2560 5.540.4 540 5.635.4 530 525 3040 3040 4383 5.840.3 2040 1285 3040 2540

Source [52,53] [53,54] [17] [36] [50] [51] [46] [47] This work [36,53] [36,53] [55,56] [57] [51,58]

Secondary Tertiary

DEA TEAe MDEA AMPf


0 (kobs

8.9 7.8 8.5 9.7

2nd/3rd 1st/2nd 1st/2nd 1st/2nd

Hindered
a b

Overall rate constant kpromoter [promoter] if rst-order with respect to the promoter or kobs kpromoter [promoter]2 if second-order with respect to the promoter). The pKa shown here is at 25 C. c The rst number represents the order of reaction with respect to the promoter, while the latter represents the overall order of reaction. d The unit of the pre-exponential factor, A, varies depending on the overall order of the reaction. If the reaction is rst order, it has the unit s1. If the reaction is second order, it has the unit M1 s1. e TEA is the abbreviation of triethanolamine. f AMP is the abbreviation of 2-amino-2-methyl-1-propanol.

278

H. Thee et al. / Chemical Engineering Journal 210 (2012) 271279 [17] G.F. Versteeg, L.A.J. Van Dijck, W.P.M. Van Swaaij, On the kinetics between CO2 and alkanolamines both in aqueous and non-aqueous solutions. An overview, Chem. Eng. Commun. 144 (1996) 113158. [18] P.M.M. Blauwhoff, G.F. Versteeg, W.P.M. Van Swaaij, A study on the reaction between CO2 and alkanolamines in aqueous solutions, Chem. Eng. Sci. 39 (1984) 207225. [19] P.V. Danckwerts, K.M. McNeil, The effects of catalysis on rates of absorption of CO2 into aqueous amine-potash solutions, Chem. Eng. Sci. 22 (1967) 925930. [20] D.W. Savage, G. Sartori, G. Astarita, Amines as rate promoters for carbon dioxide hydrolysis, Faraday Discuss. Chem. Soc. 77 (1984) 1731. [21] J.T. Cullinane, G.T. Rochelle, Kinetics of carbon dioxide absorption into aqueous potassium carbonate and piperazine, Ind. Eng. Chem. Res. 45 (2005) 25312545. [22] S.S. Laddha, P.V. Dancwerts, The absorption of CO2 by amine-potash solutions, Chem. Eng. Sci. 37 (1982) 665667. [23] M. Ahmadi, V.G. Gomes, K. Ngian, Advanced modelling in performance optimization for reactive separation in industrial CO2 removal, Sep. Purif. Technol. 63 (2008) 107115. [24] P.C. Tseng, W.S. Ho, D.W. Savage, Carbon dioxide absorption into promoted carbonate solutions, AIChE J. 34 (1988) 922931. [25] Y. Zhang, H. Chen, C.-C. Chen, J.M. Plaza, R. Dugas, G.T. Rochelle, Rate-based process modeling study of CO2 capture with aqueous monoethanolamine solution, Ind. Eng. Chem. Res. 48 (2009) 92339246. [26] H. Knuutila, O. Juliussen, H.F. Svendsen, Density and N2O solubility of sodium and potassium carbonate solutions in the temperature range 2580 C, Chem. Eng. Sci. 65 (2010) 21772182. [27] G.E.H. Joosten, P.V. Danckwerts, Solubility and diffusivity of nitrous oxide in equimolar potassium carbonate-potassium bicarbonate solutions at 25 C and 1 atm, J. Chem. Eng. Data 17 (1972) 452454. [28] G.F. Versteeg, Solubility and diffusivity of acid gases (carbon dioxide, nitrous oxide) in aqueous alkanolamine solutions, J. Chem. Eng. Data 33 (1988) 29. [29] G.A. Ratcliff, J.G. Holdcroft, Diffusivities of gases in aqueous electrolyte solutions, Trans. Inst. Chem. Eng. 41 (1963) 315319. [30] G.F. Versteeg, P.M.M. Blauwhoff, W.P.M. van Swaaij, The effect of diffusivity on gasliquid mass transfer in stirred vessels. Experiments at atmospheric and elevated pressures, Chem. Eng. Sci. 42 (1987) 11031119. [31] T. Prll, T. Todinca, M. S uta, A. Friedl, Acid gas absorption in trickle ow columnsmodelling of the residence time distribution of a pilot plant, Chem. Eng. Process. 46 (2007) 262270. [32] G. Astarita, D.W. Savage, A. Bisio, Gas Treating with Chemical Solvents, John Wiley & Sons, Inc., New York, 1983. [33] M.A. Pacheco, CO2 absorption into aqueous mixtures of diglycolamine and methyldiethanolamine, Chem. Eng. Sci. 55 (2000) 5125. [34] R.E. Treybal, Mass-Transfer Operations, McGraw-Hill, third ed., 1981. [35] D.M. Austgen, G.T. Rochelle, X. Peng, C.C. Chen, Model of vaporliquid equilibria for aqueous acid gas-alkanolamine systems using the electrolyteNRTL equation, Ind. Eng. Chem. Res. 28 (1989) 10601073. [36] H. Hikita, S. Asai, H. Ishikawa, M. Honda, The kinetics of reactions of carbon dioxide with monoethanolamine, diethanolamine and triethanolamine by a rapid mixing method, Chem. Eng. J. 13 (1977) 712. [37] R.L. Kent, B. Eisenberg, Better data for amine treating, Hydrocarbon Process. 55 (1976) 8790. [38] T.J. Edwards, G. Maurer, J. Newman, J.M. Prausnitz, Vaporliquid equilibria in multicomponent aqueous solutions of volatile weak electrolytes, AIChE J. 24 (1978) 966976. [39] B.R.W. Pinsent, L. Pearson, F.J.W. Roughton, The kinetics of combination of carbon dioxide with hydroxide ions, Trans. Faraday Soc. 52 (1956) 15121520. [40] J.I. Lee, F.D. Otto, A.E. Mather, Equilibrium between carbon dioxide and aqueous monoethanolamine solutions, J. Appl. Chem. Biotechnol. 26 (1976) 541549. [41] F.-Y. Jou, A.E. Mather, F.D. Otto, The solubility of CO in a 30 mass% monoethanolamine solution, Can. J. Chem. Eng. 73 (1995) 140147. [42] J.T. Cullinane, G.T. Rochelle, Thermodynamics of aqueous potassium carbonate, piperazine and carbon dioxide, Fluid Phase Equilibria 227 (2005) 197213. [43] N. Ramachandran, A. Aboudheir, R. Idem, P. Tontiwachwuthikul, Kinetics of the absorption of CO2 into mixed aqueous loaded solutions of monoethanolamine and methyldiethanolamine, Ind. Eng. Chem. Res. 45 (2006) 26082616. [44] A. Aboudheir, P. Tontiwachwuthikul, A. Chakma, R. Idem, Kinetics of the reactive absorption of carbon dioxide in high CO2-loaded, concentrated aqueous monoethanolamine solutions, Chem. Eng. Sci. 58 (2003) 51955210. [45] F. Leder, The absorption of CO2 into chemically reactive solutions at high temperatures, Chem. Eng. Sci. 26 (1971) 13811390. [46] J. Xiao, C.-W. Li, M.-H. Li, Kinetics of absorption of carbon dioxide into aqueous solutions of 2-amino-2-methyl-1-propanol + monoethanolamine, Chem. Eng. Sci. 55 (2000) 161175. [47] S.-Y. Horng, M.-H. Li, Kinetics of absorption of carbon dioxide into aqueous solutions of monoethanolamine + triethanolamine, Ind. Eng. Chem. Res. 41 (2001) 257266. [48] K. Onda, Mass transfer coefcients between gas and liquid phases in packed columns, J. Chem. Eng. Jpn. 1 (1968) 56. [49] K.A. Mumford, K.H. Smith, C.J. Anderson, S. Shen, W. Tao, Y.A. Suryaputradinata, A. Qader, B. Hooper, R.A. Innocenzi, S.E. Kentish, G.W. Stevens, Post-combustion capture of CO2: results from the solvent absorption capture plant at Hazelwood power station using potassium carbonate solvent, Energy Fuels 26 (2011) 138146. [50] D.E. Penny, T.J. Ritter, Kinetic study of the reaction between carbon dioxide and primary amines, J. Chem. Soc., Faraday Trans. 1: Phys. Chem. Condens. Phases 79 (1983) 21032109.

5. Conclusions A comprehensive kinetic and process modeling study on the absorption of CO2 into MEA-promoted potassium carbonate solutions has been presented in this work. Results show that the addition of MEA has signicantly accelerated the apparent pseudorst-order rate constant, and therefore, the overall absorption of CO2 into potassium carbonate solutions is improved. MEA is found to exhibit a comparable catalytic activity to that of ethylenediamine and a larger catalytic activity than that of other secondary and tertiary amines. Incorporating our experimental results into Aspen PlusTM has enabled the development of a model that can successfully simulate both industrial and pilot plant solvent capture processes employing MEA and K2CO3 as the capture solvent. This is a vital rst step to simulating MEA-promoted potassium carbonate processes. The rate constants and process model presented in this study have important implications for designing large scale absorption and regeneration units for carbon capture processes employing MEA-promoted potassium carbonate solvents.

Acknowledgments The authors acknowledge the nancial support provided by the Australian Government through its Cooperative Research Centre program for this CO2CRC research project. Infrastructure support from the Particulate Fluids Processing Centre (PFPC), a special research centre of the Australian Research Council is also gratefully acknowledged.

References
[1] K.S. Lackner, A guide to CO2 sequestration, Science 300 (2003) 16771678. [2] C. Anderson, C. Scholes, A. Lee, K. Smith, S. Kentish, G. Stevens, P.A. Webley, A. Qader, B. Hooper, Novel pre-combustion capture technologies in action-results of the CO2CRC/HRL mulgrave capture project, Energy Procedia 4 (2011) 11921198. [3] K. Smith, U. Ghosh, A. Khan, M. Simioni, K. Endo, X. Zhao, S. Kentish, A. Qader, B. Hooper, G. Stevens, Recent Developments in Solvent Absorption Technologies at the CO2CRC in Australia, ScienceDirect, 2008. [4] A.B. Rao, E.S. Rubin, A technical, economic, and environmental assessment of amine-based CO2 capture technology for power plant greenhouse gas control, Environ. Sci. Technol. 36 (2002) 44674475. [5] M.R.M. Abu-Zahra, L.H.J. Schneiders, J.P.M. Niederer, P.H.M. Feron, G.F. Versteeg, CO2 capture from power plants: Part I. A parametric study of the technical performance based on monoethanolamine, Int. J. Greenhouse Gas Control 1 (2007) 3746. [6] A. Qader, B. Hooper, T. Innocenzi, G. Stevens, S. Kentish, C. Scholes, K. Mumford, K. Smith, P.A. Webley, J. Zhang, Novel post-combustion capture technologies on a lignite red power plant results of the CO2CRC/H3 capture project, Energy Procedia 4 (2011) 16681675. [7] A.L. Kohl, F.C. Riesenfeld, Alkaline salt solutions for hydrogen sulde and carbon dioxide absorption, in: Gas Purication, Gulf Publishing Company, Houston, Texas, 1985, pp. 211246. [8] H.E. Benson, J.H. Field, R.M. Jimeson, CO2 absorption: employing hot potassium carbonate solutions, Chem. Eng. Prog. 50 (1954) 356364. [9] H.E. Benson, J.H. Field, W.P. Haynes, Improved process for carbon dioxide absorption uses hot carbonate solutions, Chem. Eng. Prog. 52 (1956) 433438. [10] J.S. Tosh, J.H. Field, H.E. Benson, W.P. Haynes, Equilibrium Study of the System Potassium Carbonate Potassium Bicarbonate, Carbon Dioxide and Water, Bureau of Mines Report of Investigations, vol. 5484, 1959, p. 23. [11] H. Thee, K.H. Smith, G. da Silva, S.E. Kentish, G.W. Stevens, Carbon dioxide absorption into unpromoted and borate-catalyzed potassium carbonate solutions, Chem. Eng. J. 181182 (2012) 694701. [12] G. Astarita, D.W. Savage, J.M. Longo, Promotion of carbon dioxide mass transfer in carbonate solutions, Chem. Eng. Sci. 36 (1981) 581588. [13] D. Guo, H. Thee, G. da Silva, J. Chen, W. Fei, S. Kentish, G.W. Stevens, Boratecatalyzed carbon dioxide hydration via the carbonic anhydrase mechanism, Environ. Sci. Technol. 45 (2011) 48024807. [14] K. Endo, Q.S. Nguyen, S.E. Kentish, G.W. Stevens, The effect of boric acid on the vapour liquid equilibrium of aqueous potassium carbonate, Fluid Phase Equilib. 309 (2011) 109113. [15] G.S. Goff, G.T. Rochelle, Monoethanolamine degradation: O2 mass transfer effects under CO2 capture conditions, Ind. Eng. Chem. Res. 43 (2004) 6400 6408. [16] P.V. Danckwerts, The reaction of CO2 with ethanolamines, Chem. Eng. Sci. 34 (1979) 443446.

H. Thee et al. / Chemical Engineering Journal 210 (2012) 271279 [51] E. Alper, Reaction mechanism and kinetics of aqueous solutions of 2-amino2-methyl-1-propanol and carbon dioxide, Ind. Eng. Chem. Res. 29 (1990) 17251728. [52] S. Bishnoi, G.T. Rochelle, Absorption of carbon dioxide into aqueous piperazine: reaction kinetics, mass transfer and solubility, Chem. Eng. Sci. 55 (2000) 55315543. [53] R.N. Goldberg, Thermodynamic quantities for the ionization reactions of buffers, J. Phys. Chem. Ref. Data 31 (2002) 231. [54] H. Hikita, S. Asai, H. Ishikawa, M. Honda, The kinetics of reactions of carbon dioxide with monoisopropanolamine, diglycolamine and ethylenediamine by a rapid mixing method, Chem. Eng. J. 14 (1977) 2730.

279

[55] G.F. Versteeg, W.P.M. van Swaaij, On the kinetics between CO2 and alkanolamines both in aqueous and non-aqueous solutionsII. Tertiary amines, Chem. Eng. Sci. 43 (1988) 587591. [56] D. Barth, C. Tondre, J.J. Delpuech, Kinetics and mechanisms of the reactions of carbon dioxide with alkanolamines: a discussion concerning the cases of MDEA and DEA, Chem. Eng. Sci. 39 (1984) 17531757. [57] J.-J. Ko, M.-H. Li, Kinetics of absorption of carbon dioxide into solutions of Nmethyldiethanolamine + water, Chem. Eng. Sci. 55 (2000) 41394147. [58] G. Sartori, D.W. Savage, Sterically hindered amines for carbon dioxide removal from gases, Ind. Eng. Chem. Fundam. 22 (1983) 239249.

Vous aimerez peut-être aussi