Вы находитесь на странице: 1из 24

This article was downloaded by: [Indian Institute of Petroleum] On: 20 June 2012, At: 01:48 Publisher: Taylor

& Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Mathematical and Computer Modelling of Dynamical Systems: Methods, Tools and Applications in Engineering and Related Sciences
Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/nmcm20

CO2 removal by absorption: challenges in modelling


L.E. i
a a

Department of Technology, Telemark University College, PO Box 203, N-3901, Porsgrunn, Norway Available online: 15 Jul 2010

To cite this article: L.E. i (2010): CO2 removal by absorption: challenges in modelling, Mathematical and Computer Modelling of Dynamical Systems: Methods, Tools and Applications in Engineering and Related Sciences, 16:6, 511-533 To link to this article: http://dx.doi.org/10.1080/13873954.2010.491676

PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.tandfonline.com/page/terms-andconditions This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Mathematical and Computer Modelling of Dynamical Systems Vol. 16, No. 6, December 2010, 511533

CO2 removal by absorption: challenges in modelling


L.E. i*
Department of Technology, Telemark University College, PO Box 203, N-3901 Porsgrunn, Norway (Received 4 September 2009; final version received 5 May 2010) The traditional method for large-scale CO2 removal is by absorption in a mixture of an amine and water. The tasks of modelling this process can be divided into descriptions of absorption and reaction kinetics, gas/liquid equilibrium, gas and liquid flows and pressure drop. Process simulation tools containing models for most of these tasks are commercially available, and the calculated results can be used as a basis for equipment dimensioning and economical optimization. A flowsheet calculation in the program Aspen HYSYS is used as an example. Calculation convergence is important, especially the column convergence is critical. For some simplified conditions, calculation of stage efficiencies can give a satisfactory description of the absorption process. Computational fluid dynamics is an efficient tool for calculating flow conditions, pressure drop and temperature profiles, especially for one-fluid phase. An unsolved problem when using computational fluid dynamics for gas/liquid processes is the description of the gas/liquid interfacial area. A major challenge is to combine different models and calculation tools. An improved model for a specific task must be available and possible to combine with other calculation tools to be utilized by other programs. In an example, models for equilibrium and mass transfer efficiency are used in a flowsheet calculation including CO2 absorption and desorption, followed by economical optimization. The example illustrates some possibilities, limitations and challenges. Keywords: CO2; amine; absorption; simulation; modelling; efficiency

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

1. Introduction CO2 has been removed from process streams at an industrial scale since about 1930. The most important removal processes have been from natural gas and in the removal of CO2 from industrial gases at high pressures for ammonia and methanol production. The main process is absorption in a mixture of an amine and water. Other solvents like carbonate salt solutions have also been used. An overview of processes can be found in Kohl and Nielsen [1]. CO2 removal from exhaust gases has received much interest because of the environmental need for reducing CO2 emissions to the atmosphere. Many processes for removal of CO2 from power plants or other exhaust gases have been suggested. The emphasis here is put on post-combustion absorption in an amine-based solvent. This is the most relevant method for gas-based power plants in the near future. The technology is relatively mature and can be utilized by existing plants with little modification. Absorption is traditionally performed in a column with plates, random packing or structured packing; CO2-containing gas flows upwards and the absorption liquid flows downwards. The absorbed CO2 is regenerated in a desorption column, and the solvent is recirculated to the absorption
*Email: lars.oi@hit.no
ISSN 1387-3954 print/ISSN 1744-5051 online 2010 Taylor & Francis DOI: 10.1080/13873954.2010.491676 http://www.informaworld.com

512
Purified gas

L.E. i

Product CO2 Condenser Amine cooler Desorber Amine/amine exchanger

CO2 absorber Exhaust gas

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

Reboiler

Figure 1.

A general CO2 absorption and desorption process.

column. A removal process consisting of absorption, desorption, heat exchangers and auxiliary equipment is shown in Figure 1. The main challenges in modelling the CO2 removal are in the absorption and desorption processes. The absorption column is the largest and most expensive unit, and the most important chemical reactions take place in this column. The main energy consumption is in the reboiler connected to the desorption column. Equipment units like heat exchangers and pumps in the CO2 removal process are little different from similar equipment in other chemical processes. In CO2 absorption and desorption, the modelling challenges can be divided into the following tasks:     Absorption and reaction kinetics Gas/liquid equilibrium Gas and liquid flows Pressure drop

Calculation methods for most of these tasks have been available for a long time. Danckwerts and Sharma [2] wrote a review as early as in 1966. When computers were introduced for chemical engineering calculations, computer programs were made to perform these calculations. The calculations can also be extended to comprise tools for, for example, mechanical equipment dimensioning, cost estimation and economical optimization. Several commercial process simulation programs have included models especially for the amine/water/CO2 system. Examples of such programs are Aspen HYSYS, Aspen Plus, Pro/II and ProMax. These programs have built-in models especially for vapour/ liquid equilibrium and calculation tools especially for column solving. Some of them also have kinetic models available. Most of the tasks in the list above can be handled in a process simulation program. Computational fluid dynamics (CFD) modelling can be used to calculate the flow phenomena in absorption and desorption columns. This can then be used to calculate liquid holdup, pressure drop and capacity. Fluent and CFX are examples of commercial CFD programs. An optimistic aim for CFD modelling of absorption and desorption is to contribute to a complete, detailed and quantitative description of the absorption/desorption process.

Mathematical and Computer Modelling of Dynamical Systems

513

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

This article first gives an overview of process simulation tools that have been used for CO2 removal. Then some details of the chemistry, kinetics and equilibrium for the process are presented. An overview of models describing the mechanisms of absorption combined with chemical reaction is presented. Traditional design methods for calculating pressure drop, interfacial area and mass transfer are presented, followed by similar calculation methods using CFD. The last part of this article discusses the combination of different models. In an example, a flowsheet calculation is performed which combines models for equilibrium and mass transfer efficiency with calculation tools for column solving and cost optimization. This example illustrates some possibilities, limitations and challenges. The models for solving specific tasks of the CO2 absorption process differ in accuracy, complexity and robustness. The main purpose of this article is to give an overview and to pinpoint areas where further modelling can lead to important improvements. A major challenge is to combine different models and calculation tools.

2. Flowsheet calculation of CO2 removal using process simulation tools 2.1. Use of process simulation programs for absorption and desorption Process simulation programs such as Aspen HYSYS, Aspen Plus, Pro/II and ProMax have been used to calculate CO2 removal by absorption, mainly at steady-state conditions. The main advantages of these process simulation programs are that a large number of models for vapour/liquid equilibrium and also different calculation tools for unit operations are available. Simulation of CO2 removal from flue gas in a monoethanol amine (MEA)/water system has been performed by Desideri and Paolucci [3] and by Alie et al. [4]. Both have used the simulation program Aspen Plus with the MEA property insert, which is based on the Chen/ Austgen electrolyte-non-random two-liquid (NRTL) equilibrium model [5,6]. Desideri and Paolucci used a specified number of theoretical stages in the absorption and desorption column. Tobiesen et al. [7] and Aroonwilas et al. [8] have made Fortran programs to perform similar calculations. All of the above-mentioned references calculate steady-state solutions. Kvamsdal et al. [9] have simulated the absorption part of the process dynamically (as a function of time) using gPROMS as a modelling tool. At Telemark University College, Matlab has also been used to model CO2 absorption and desorption dynamically [10]. An important feature of commercial process simulation programs is the available tools for achieving convergence of the process flowsheet (containing unit operations and streams). The early process simulation programs (Pro/II and Aspen Plus) were sequential modular and had to be solved in a forward manner. A traditional way to calculate the flowsheet is to use recycle convergence blocks to compare updated streams with earlier calculated streams. Some simulation programs (like Aspen HYSYS) can calculate the flowsheet backwards, but because of limitations especially in the column units, flowsheet convergence is normally achieved as if the program was sequential modular. For flowsheet convergence, the programs can make use of recycle blocks, nested or simultaneous calculation sequences and different acceleration methods like Wegstein or dominant eigenvalue methods. Convergence of the columns is particularly important and was the core technology in the early steady-state process simulation programs. Because the CO2/amine/water system is highly non-ideal, the need for efficient and robust column solvers is of major importance. Column solver methods available in addition to the default Hysim Inside-Out method in the Aspen HYSYS program are

514
Make Q-Lean up cooler water Lean MEA recycle Make up MEA

L.E. i

Recycle RCY-1

Lean MEA to lean cooler Lean cooler

MIX-100

Sweet gas Lean MEA

Rich MEA to rich/lean heat exchanger

Rich/lean heat exchanger Q-Condenser Rich MEA to desorber Q-Reboiler

CO2

Rich pump

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

Sour gas Absorber

Rich MEA Q-Rich pimp Lean MEA to rich/lean heat exchanger

Lean pump

Desorber

SPRDSHT-1

Q-Lean pump

Lean MEA from desorber

Figure 2.

Aspen HYSYS flowsheet for a CO2 removal process (from i [11]).

     

Modified Hysim Inside-Out Newton Raphson Inside-Out Sparse Continuation Solver Simultaneous Correction OLI Solver Fixed or adaptive damping factor

An Aspen HYSYS flowsheet from i [11] is shown in Figure 2. Specifications for the calculation are given in Table 1. The Aspen HYSYS version 2004.2 was used with the Kent Eisenberg equilibrium model [12], which is extended in Aspen HYSYS by Li and Shen [13]. If there are too many stages specified in the columns, they tend to diverge, which is traditional for column stage calculations. It was found that the Modified Hysim Inside-Out method with adaptive damping gives the best convergence [11]. The base case calculated a CO2 removal rate of 85% with a heat consumption of 3.67 MJ/ kg CO2. Figure 3 shows the calculated CO2 removal efficiency and heat consumption when the amine circulation rate is varied in the Aspen HYSYS flowsheet calculation. The calculation shows a minimum heat consumption at a certain circulation rate. 2.2. Rigorous simulation in process simulation programs

Most of the column models in commercial process simulation programs are based on equilibrium stages or stages with a stage efficiency. More rigorous column models, which include kinetic expressions, are available. Some of these are able to calculate the concentration profiles of all the diffusing components through the liquid film near the gas/liquid surface. This kind of approach is based on solving the differential equations describing the diffusion and chemical kinetics in the liquid film. An approach for rigorous modelling with resulting differential equations is presented in Section 5.5.

Mathematical and Computer Modelling of Dynamical Systems


Table 1. Specifications for base case CO2 removal.

515

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

Inlet gas temperature Inlet gas pressure Inlet gas flow CO2 in inlet gas Water in inlet gas Lean amine temperature Lean amine pressure Lean amine rate MEA content in lean amine CO2 in lean amine Number of stages in absorber Murphree efficiency in absorber Rich amine pump pressure Heated rich amine temperature Number of stages in stripper Murphree efficiency in stripper Reflux ratio in stripper Reboiler temperature Lean amine pump pressure Minimum delta T in heat exchange
Note: aIn first iteration.

40 C 1.1 bar (abs) 85,000 kmol/h 3.73 mol% 6.71 mol% 40 C 1.1 bar (abs) 120,000 kmol/ha 29 mass%a 5.5 mass%a 10 0.25 2 bar 104.5 Ca 6 (3 + 3) 1.0 0.3 120 C 2 bar 10 C

90

4.2 (*) Heat consumption (MJ/kg CO2)

(0) CO2 removal (%)

85

80

3.8

75

2.5

2.6 2.7 2.8 2.9 3 Circulation rate (1000 ton/h)

3.1

3.6 3.2

Figure 3. i [11]).

CO2 removal grade and heat consumption as a function of amine circulation rate (from

The program Aspen Plus has possibilities to include such calculations, contrary to, for example, Aspen HYSYS. Al-Baghli et al. [14] have made a rate-based model for the design of gas absorbers for the removal of CO2 and H2S using aqueous solutions of MEA and diethanol amine (DEA). Freguia and Rochelle [15] used a Fortran subroutine integrated into Aspen Plus to perform a rate-based calculation of CO2 absorption into MEA. Kucka et al. [16] have used the Aspen Custom Modeler tool in Aspen Plus to model the liquid film by dividing the film into a number of segments. The mentioned examples in Aspen Plus have been performed at steady state.

516 2.3.

L.E. i Difficulties and challenges using process simulation tools for CO2 absorption

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

Dynamic simulations are of interest for regulation, for start-up and shut-down analysis and for evaluation of emergency situations. Most of the process simulation calculations of CO2 removal from atmospheric exhaust in the open literature are simplified processes at steady state. The unit operation modules in the process simulation programs Aspen HYSYS, Aspen Plus and Pro/II are traditionally based on the solving of algebraic equations for a steady-state solution. All these programs have the possibility to calculate a flowsheet as a function of time by performing time steps. There are many challenges met when trying to extend steady-state simulations to dynamic conditions. Reducing computer time and increasing calculation robustness are two challenges. There are of course challenges to improve the different models in a process simulation program. There is perhaps a more important challenge on how to combine the different models. Because process simulation programs utilize many rigorous models, calculation convergence can be a problem. Alie et al. [4] use a flowsheet decomposition approach, dividing the calculation of the absorption column and desorption column, to overcome this problem. Detailed absorption kinetics models can be complex and can involve nonlinear equilibrium models and the solving of partial differential equations. It is doubtful whether such calculations should be performed for each iteration when a flowsheet involving absorption and desorption columns is calculated several times. 3. Chemistry of CO2 removal with amines 3.1. Primary, secondary and tertiary amines A general amine has the formula NR1R2R3 where R1, R2 and R3 are organic groups or hydrogen directly bonded to the central nitrogen atom. An amine with only one organic group directly bonded to nitrogen is a primary amine, with two organic groups it is a secondary amine and with three it is a tertiary amine. If an organic group contains an OH group, the amine is called an alkanolamine. MEA (H2NC2H4OH) is a primary alkanolamine often used for CO2 removal. DEA is a simple secondary amine and N-methyldiethanolamine (MDEA), with R1 and R2 as C2H4OH-groups and R3 as a CH3-group, is probably the most used tertiary amine for CO2 removal. When used as solvents, the amines are typically 2040 wt.% solutions in water. 3.2. Absorption of CO2 into amine solutions including chemical reactions The absorption of CO2 into an amine solution with MEA can be described by the following equations: Equation (1) describes the transfer of CO2 from gas to an aqueous liquid, Equation (2) describes the reaction to a protonated amine ion (HMEA+) and a carbamate ion, and bicarbonate HCO formation according to Equation (3) is also occurring. (A 3 carbamate ion is a reaction product formed by CO2 and an amine, and if the amine is MEA, the carbamate ion has the formula HN(C2H4OH)COO-.) In the case of other amines than MEA, a reaction equivalent to Equation (3) can be more important than the reaction in Equation (2). CO2 g $ CO2 aq CO2 2MEA $ HMEA Carbamate (1) (2)

Mathematical and Computer Modelling of Dynamical Systems CO2 MEA H2 O $ HMEA HCO 3

517 (3)

CO2 in an aqueous solution can be in the form of free (or molecular) CO2 or as bicarbonate or carbonate ions (HCO or CO2 ). When an amine is added, carbamate can 3 3 be formed according to Equation (2). The total concentration of CO2 (at a molar basis) is the sum of all the concentrations of the different forms: CCO2 ;TOT CCO2 CHCO CCO2 CCARBAMATE 3 3 (4)

The amine MEA can be in the form of free MEA, protonated MEA (HMEA+) or as a part of a carbamate, The total concentration of MEA is the sum of the concentrations of the different forms:

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

CMEA;TOT CMEA CHMEA CCARBAMATE

(5)

3.3. Reaction kinetics for MEA (and other primary and secondary amines) with CO2 The detailed reaction kinetics for the reaction between even simple amines and CO2 are quite complicated. However, for the simple amines, for example MEA, the kinetics are now regarded as well known. Overviews can be found in the review articles by Danckwerts and Sharma [2] and Versteeg et al. [17]. The mass transfer kinetics of MEA absorption in laboratory absorption equipment under controlled conditions can be explained by traditional mass transfer models [17]. It has been known for a long time that the main reaction described by Equation (2) of primary amines like MEA and CO2 is a second-order reaction under normal conditions: rCO2 k2 CCO2 CMEA (6)

In Equation (6), rCO2 is the reaction rate (in mole CO2 reacted per volume and time), k2 is the reaction rate constant (which is temperature dependent) and CCO2 and CMEA are the concentrations of free CO2 and MEA. Versteeg et al. [17] give temperature-dependent values for the reaction rate constant. Caplow [18] gives a detailed description of the reaction kinetics, introducing the so-called zwitterion mechanism. This has later been generally accepted as the actual mechanism [17]. The mechanism is similar for simple primary and secondary amines.

3.4.

Kinetics for MDEA (and other tertiary amines) and sterically hindered amines

MDEA is a tertiary amine and does not react with CO2 according to the carbamate formation reaction in Equation (2). The absorption of CO2 is in this case instead followed by an acid/ base reaction as in Equation (3). The reaction is described in Blauwhoff et al. [19] and Rinker et al. [20]. The mechanism is similar for many of the tertiary amines. Because tertiary amines do not form carbamates, they normally have lower desorption energy compared to primary amines. Not all primary and secondary amines react with CO2 to form carbamate. Because of bulky groups close to the nitrogen atom in the amine group, some primary and secondary

518

L.E. i

amines do not react with CO2 to form carbamates. These are so-called sterically hindered amines. One example is 2-amino-2-methyl-1-propanol (AMP). The sterically hindered amines perform in contact with water and CO2 in many ways like tertiary amines. The idea of using sterically hindered amines is described by Sartori and Savage [21]. The solvent KS-1 (which is based on sterically hindered amines) is used by Mitsubishi Heavy Industries in their commercial process for CO2 removal from flue gases. The KS-1 process is claimed to have much lower energy consumption than an MEA-based process [22]. 3.5. Reaction kinetics for mixtures of amines with CO2 Using mixtures of amines for CO2 removal is first described by Chakrawarty et al. [23]. One idea is to combine the high reactivity of one amine (e.g. MEA) with the low desorption energy of another amine (e.g. MDEA). The reaction kinetics can normally be described by the kinetics of the single amines. The combination of reaction kinetics with mass transfer will however be much more complicated in the case of mixed amines. The combination of reaction kinetics, diffusion and equilibrium is treated in more detail in Section 5. 4. Equilibrium description of mixtures of water, CO2 and amines 4.1. Gas/liquid equilibrium Equilibrium between the CO2 content in the gas and the concentration of free CO2 in the liquid phase can be described by a Henrys constant (He), which is temperature dependent: pCO2 He CCO2 (7)

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

pCO2 is the partial pressure of CO2 and CCO2 is the concentration (on a molar basis) of free CO2 in the liquid. Under atmospheric conditions, gas non-idealities are normally negligible, and the ideal gas law is sufficient to describe the gas phase. An equation of state like Peng Robinson [24] can also be used to take care of the minor gas non-idealities. In that case, the partial pressure of CO2 is replaced by the fugacity of CO2. A water/amine/CO2 mixture can be specified by the total concentrations of CO2 and amine. The gas/liquid equilibrium can be expressed by the equilibrium between the partial pressures of CO2 in the gas above a specified solution at a given temperature as indicated in Equation (8): pCO2 f CCO2 ;TOT ; CAMINE;TOT ; T (8)

Experimental gas/liquid equilibrium data for amine systems are often measured as partial pressures of CO2 in the gas above a specified solution, and the data will then be in the form of Equation (8). 4.2. General chemical equilibrium models Simple models which empirically fit measured data, for example, according to Equation (8) are available. Such models will, however, not give any information about the actual composition of the liquid. Kent and Eisenberg [12] gave an equilibrium description based on literature values for Henrys constants and equilibrium constants for the water/carbonate/bicarbonate system. The water, carbonate and bicarbonate system is a widely studied and well-described system [2].

Mathematical and Computer Modelling of Dynamical Systems

519

Then the equilibrium constants for the amine/carbamate equilibrium and the amine/protonated amine (Equations (2) and (3)) were fitted to experimental data. A modified version of this model, Li/Shen [13], is used by the simulation program Aspen HYSYS. 4.3. Activity-based equations/electrolyte models

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

A more detailed description can be done by expressing the activities (or chemical potentials) of all the ionic and molecular components as a function of liquid concentrations and temperature. The Chen/Austgen model [6], for simple amine systems, is based on the general electrolyte-NRTL model of Chen and Evans [5]. This model is available in the simulation program Aspen Plus. This is a rigorous model and has a high complexity, many adjustable parameters and high accuracy. Liu et al. [25] have adjusted the parameters for the MEA/ water system from the Chen/Austgen model to make the heat of vapourization to be more accurate. Li/Mather [26] is a similar model available in Aspen HYSYS, using an electrolyteMargules model. Aspen HYSYS is in principle not a program suitable for describing electrolytic systems. It does not calculate the resulting concentrations of ionic species in the solution. When the Li/Mather model is calculated in Aspen HYSYS, the result is the concentrations of the constituting components of the solution (e.g. CCO2 TOT and CMEA,TOT). The calculation based on ionic species is kept inside the subroutine containing the model. Kaewsichan et al. [27] gave an overview over equilibrium models in amine systems and also presented a model based on an electrolyte UNIversal-QUAsi-Chemical (UNIQUAC) model. An extended UNIQUAC model has also been used to describe the CO2/amine system at Technical University of Denmark [28]. The statistical associating fluid theory variable range (SAFT-VR) [29] has been used by Imperial College [30] to model the equilibrium and kinetics in amine/water/CO2 systems. This has been implemented in the program package gPROMS. The use of such a molecular approach is especially interesting for modelling kinetic and equilibrium properties for new solvents with limited available equilibrium data. The accuracy of models for amine mixtures is often limited by the accuracy in the available equilibrium data. Equilibrium models often have a trade-off between a complex model with high accuracy and a simpler model with less accuracy. There is a challenge to find equilibrium models that are simple, accurate and achieve convergence easily. 5. Models for absorption of CO2 followed by chemical reaction 5.1. Description of the absorption process followed by chemical reaction A schematic overview of typical partial pressure and concentration profiles at a certain column height in the absorber is given in Figure 4. The figure is based on the two-film concept. Absorption of a gas is described by transport from the bulk gas to the liquid surface, the assumption of gas/liquid equilibrium at the interface and then transport of absorbed gas to the liquid bulk. After absorption, CO2 can either react directly in the liquid close to the interface or be transported into the bulk liquid. In the bulk liquid, CO2 or other species can react further, limited by either equilibrium or chemical kinetics. The concentration of amine decreases from the bulk to the liquid film as shown in Figure 4 because it reacts with CO2 mainly in the film. Amine also evaporates to some extent to the gas phase. This is, however, not important for the CO2 absorption and reaction mechanisms. The rate of amine evaporation is important for the study of amine emissions from the absorption column and the possibility to reduce amine emissions to the atmosphere.

520
Concentration or partial pressure (mol/m3) or (Pa) Gas bulk Gas film PCO , Interface
2

L.E. i
Liquid film Liquid bulk

PCO2

CCO2, Interface

CAmine

CCO2

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

Distance (m)

Figure 4. Typical concentration profiles in liquid film with absorption and chemical reaction, assuming equilibrium between partial pressure and concentration of CO2 at interface.

5.2. Mass transfer models Popular mass transfer models are the two-film theory by Lewis and Whitman [31] and the penetration model and surface renewal model developed by Higbie [32] and Danckwerts [33], respectively. The two-film model is based on the concept of thin gas and liquid films with a constant thickness and transport rate based on molecular diffusion. This model results in mass transfer proportional to diffusivity DCO2 . The penetration and the surface renewal models are regarded to be more realistic models for the liquid film. They are based on the idea of continuous transport of volume elements from the interface to the liquid bulk. The difference between the models is that the penetration model assumes equal contact time for the elements, whereas the surface renewal model assumes a contact time distribution. Both of these models result in mass transfer proportional to the square root of the diffusivity. The boundary layer theory has also been used to calculate mass transfer from basic laws of fluid dynamics, mainly for simple geometries [34]. A mass transfer number for the liquid film, kL, can be defined by Equation (9). The gas/ liquid interfacial area per volume has symbol a, and the interface and bulk positions are shown in Figure 4. RCO2 is used as the symbol for the absorption rate of CO2 per total volume. The reaction rate rCO2 of CO2 in Equation (4) is per liquid volume. At steady-state conditions, and assuming that the amount of unreacted CO2 is negligible compared to the reacted CO2, the rate of reaction equals the rate of absorption for CO2: RCO2 NCO2 a kL a CCO2 ;INTERFACE CCO2 ;BULK : (9)

The gas transport of CO2 to the interface is normally not rate-limiting [2]. The gas side mass transfer can often be neglected or it can be described by a simple empirical correlation. 5.3. Simplified models for absorption followed by chemical reaction These kinds of processes have been treated by, for example, Van Krevelen and Hoftijzer [35]. They used an enhancement factor which is the ratio of the actual absorption rate divided

Mathematical and Computer Modelling of Dynamical Systems

521

by the absorption rate by pure mass transfer. This can be expressed as the ratio between the mass transfer number with and without reaction: EENH kL kL;WITHOUT
REACTION

(10)

In the case where the rate expression can be assumed to be independent of the amine concentration, the pseudo-first-order conditions occur. In that case, the rate expression becomes [35] p RCO2 CCO2 ; INTERFACE a k2 DCO2 CMEA (11) where DCO2 is the molecular diffusivity of CO2 in the liquid. With this rate expression, it is possible to calculate a Murphree efficiency for a tray or a packed section at the given flow conditions. The Murphree efficiency for a tray or a packing section can be defined as shown in Figure 5. Aspen HYSYS has a model for calculating Murphree efficiencies in plate columns based on a pseudo-first-order expression. This model is based on the work by Tomkej et al. [36]. For the conditions in Table 1, Aspen HYSYS calculates a Murphree efficiency that varies slightly from top to bottom with an average value of approximately 0.09. If a plate distance of 0.6 m is assumed, this is equivalent to a Murphree efficiency of 0.15/m of packing. The Murphree efficiency increases slightly with temperature. There has been much work on calculation methods for enhancement factors and evaluations of the conditions for when to use approximate expressions. DeCoursey [37] developed

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

y y* Tray n (or section n)

yn1 Tray n1

Figure 5. Definition of Murphree efficiency, EMURPHREE (y - yn-1)/(y* - yn-1), where y is mole fraction (CO2) in the gas phase leaving tray (or section) n, and y* is in equilibrium with the liquid on tray (or section) n.

522

L.E. i

an explicit approximate expression for the case of absorption followed by a second-order irreversible reaction. Later, this was extended to reversible reactions [38], like CO2 absorption in an amine solution. Tobiesen and Svendsen [39] have calculated enhancement factors for CO2 absorption into MEA for both simple models and a rigorous model and compared it with pilot scale experiments. In a temperature range between 50 and 70 C, the deviation using a pseudo-first-order approximation was in the order of magnitude area between 5% and 50%. The deviation increased with CO2 concentration and temperature. For the case of CO2 absorption into MEA solutions at atmospheric conditions, it is not clear under which conditions the pseudo-first-order approximation is valid. 5.4. Models for CO2 absorption into mixed amines

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

In the case of mixed amines, the combined reaction and mass transfer kinetics might be quite complicated. One model is the shuttle mechanism from Astarita et al. [40], which tries to model absorption into a mixture of a reactive amine (e.g. MEA) in small amounts and a less reactive amine (e.g. MDEA) in larger amounts. The idea is that CO2 first reacts with MEA in the film close to the surface and is transferred into the bulk liquid. In the bulk liquid, CO2 is released from MEA and reacts with MDEA, so that the MEA can be shuttled back to the liquid film. Hagewiesche et al. [41] have modelled absorption into blends of MEA and MDEA. The absorption mechanism was shown to follow the shuttle mechanism proposed by Astarita. 5.5. Rigorous simulation

There have been several attempts to calculate the concentration profiles through the liquid films based on available mass transfer and kinetic models. De Leye and Froment [42], AlBaghli et al. [14] and Kucka et al. [16] are examples. Equations (12) and (13) are from DeCoursey [37] for the case of a second-order irreversible reaction between an absorbed component (e.g. CO2) and a liquid component (e.g. MEA). Mass transfer is based on a surface renewal model [33]. The equations represent a time-dependent material balance for CO2 and MEA: DCO2 @ 2 CCO2 @CCO2 k2 CCO2 CMEA 0 @x2 @t (12)

DMEA

@ 2 CMEA @CMEA 2 k2 CCO2 CMEA 0: @x2 @t

(13)

The boundary conditions are that for t 0 and x > 0 and for t > 0 and x 1, CCO2 and CMEA are equal to the bulk concentrations; and for t > 0 and x 0, CCO2 is the interface concentration and CMEA/x 0. The boundary condition that CMEA/x 0 at the interface is based on no transport of amine into the gas [37]. The solution of these equations gives the concentration profiles through the liquid film at a given time t. The accuracy is of course limited to the accuracy of the data used and the assumptions taken. Such models can be applied in steady-state flowsheet calculations if a contact time or contact time distribution (as in the surface renewal model) is assumed. In that case, the flux NCO2 at the gas/liquid interface is calculated by Equation (14) from the solution of the differential equations. The average absorption rate per volume is then calculated as the mean flux as shown by Equation (15):

Mathematical and Computer Modelling of Dynamical Systems  NCO2 DCO2 @CCO2 @x Z NCO2 dt:
0

523 (14)


x0

RCO2 N CO2

a a

(15)

5.6.

Rigorous simulation of reversible reactions

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

For the case of reversible reactions, the description of the reactions becomes more complicated. One new equation is necessary for each relevant reaction product, the rate expression in Equation (6) must be extended and the equilibrium must be taken into consideration. A simple way to describe the equilibrium is to specify an equilibrium constant based on concentrations. There is a challenge to combine rigorous kinetic models with rigorous equilibrium models. The most accurate equilibrium models are based on activities. A consistency problem may occur if the kinetic expressions used are based on concentrations and the equilibrium expressions are based on activities. 6. Pressure drop, interfacial area, mass transfer, gas and liquid distribution in columns 6.1. Traditional design methods for random and structured packing Design of packed columns is generally based on empirical correlations for liquid hold-up, pressure drop, gas/liquid interfacial area and mass transfer. The resistance to absorption is often divided into gas side and liquid side resistance. These methods are described in, for example, Kohl and Nielsen [1]. Structured packing columns will probably be the primary choice in case of a large-scale CO2 removal process from atmospheric exhaust. Structured packing is very efficient and gives a very low pressure drop. Plate columns will probably not be practical for columns with large diameters (more than 15 m). Large plates will need extensive mechanical support, and horizontally flowing liquid will need very long flow paths for each plate. Random packing will have lower investment than structured packing and might be an economical alternative. The gas flow (or gas capacity) in an absorption column is limited by pressure drop, loading or flooding. The loading point can be defined as the point where mass transfer efficiency drops significantly, and the flooding point can be defined as the condition of restricted liquid downward flow leading to liquid filling of the column. To calculate flooding (capacity) and pressure drop in random packing, empirical charts or equations as in Sherwood et al. [43] and Eckert [44] are traditional methods. They are based on correlations from dimensional analysis which are fitted to performance data. The empirical Onda [45] and Bravo and Fair [46] correlations are standard methods to calculate mass transfer in random packing. They have different correlations for calculating the gas side and liquid side mass transfer. Design methods for structured packing are based on the same type of correlations as for random packing, for example Rocha et al. [47,48], Billet and Schultes [49] and De Brito et al. [50]. Most of these methods are limited to the flow regime below the loading point. Droplet formation (which occurs above the loading point), and its influence on

524

L.E. i

interfacial area and mass transfer, is difficult to predict. Review articles for mass transfer in structured packing are written by Brunazzi et al. [51], Valluri et al. [52] and Wang et al. [53]. The semi-empirical calculation methods for mass transfer are traditionally based on the following calculation steps [51]:     liquid hold-up gas/liquid interfacial area mass transfer coefficient for gas side mass transfer coefficient for liquid side

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

The deviation between the estimation methods is especially large for the calculated effective interfacial area. This is an important parameter because the absorption rate is normally proportional to this entity, for example as in Equation (11). There is a large potential in improving the estimation methods for the effective interfacial area. 6.2. Non-empirical modelling of absorption in structured packing In their review, Valluri et al. [52] have one section for non-empirical modelling of the design parameters. They refer to Shetti and Cerro [54] as the first complete model of this kind. One of their aims was to estimate design parameters in structured packing without any adjustable parameters. An idea was to establish the equations for the fluid flow pattern and mass transfer through the films and then solve the equations to achieve the design parameters for heat and mass transfer. The equations to be solved are typically a set of algebraic and differential equations. Another early presentation of a mechanistic model for mass transfer in structured packing is by Nawrocki et al. [55]. Figure 6 is an illustration of important factors in modelling flow in structured packings. The liquid flows downwards, normally covering most of the solid surface area. The liquid velocity is largest close to the gas/liquid interface area. The gas flows upwards in the space not occupied by solid and liquid. The gas velocity is lowest close to the liquid (or solid) surface. In some models, it is assumed that the liquid (and possibly also the gas) is perfectly mixed at certain mixing points. At Delft University, models for columns with structured packing have been studied. Olujic et al. [56,57] distinguish between modelling at a geometric macro-level (channel

Liquid flow Solid sheets (possibly with corrugation and holes) Solid surface

Gas flow

Mixing points

Gas/liquid interfacial area

Figure 6.

Illustration of important factors in modelling flow in structured packing.

Mathematical and Computer Modelling of Dynamical Systems

525

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

dimensions) and micro-level (film and surface texture dimensions). Models for film flow, gas side mass transfer and liquid side mass transfer are suggested. Their prediction method does not require packing specific constants. It is stated that a reliable prediction of the effective interfacial area is the key to the success of a prediction method. Shilkin and Kenig [58], from the University of Dortmund, have made a model for structured packing columns giving a set of differential equations. The concept is based on two phases which are totally mixed at regular intervals as indicated in Figure 6. The equations are solved numerically. The results are the velocity profiles, the concentration profiles and the temperature profiles through the column. Iliuta and Larachi [59], at the Laval University in Quebec, have made a mechanistic model for structured packing columns, calculating pressure drop, liquid hold-up and wetted area. The model is based on a double-slit mechanistic approach. In a channel, the liquid film flows downwards in one slit and the gas upwards in another slit. The resulting model gives three coupled algebraic equations to be solved. The model requires no adjustable parameters. Their work has been developed further into CFD modelling. 6.3. CFD modelling of separation columns with structured packing A CFD program divides a fluid flow geometry into a grid of small volumes and then solves the fundamental equations for mass, energy and momentum conservation for each volume. Modelling of turbulence is an important part of a CFD program. Equations for chemical kinetics and equilibrium can be included. Because a CFD simulation consists of a large number of equations, CFD simulation consumes much computer memory and time. Fluent and CFX are commercial CFD programs. Valluri et al. [52] state that very few publications have been presented in the field of using CFD for structured packings. Most of them are about catalytic reactors. However, mass transfer both in gas and liquid in structured packing was also covered. Klker et al. [60] have tried to integrate CFD and process simulation for reactive distillation in structured packing. Petre et al. [61], from the Laval group in Quebec, calculated dry pressure drop in structured packing for large-scale absorption with 3D CFD. The CFD program Fluent was used with the ReNormalized-Group (RNG) k-" turbulence model. Larachi et al. [62] and Iliuta et al. [63] calculated the pressure drop for two-phase flow using CFD. The types of structured packing studied were MellaPak, GemPak, Sulzer BX and Montz-Pak. Raynal et al. [64] wrote an article called Liquid holdup and pressure drop determination in structured packing with CFD simulations. Dry pressure drop was calculated in 3D CFD using Fluent with the k-" turbulence model and the RNG k-" model. Hold-up was calculated using a 2D laminar model. The calculations were compared with experiments from an air/water system. Raynal et al. [65] have also written an article called Use of CFD for CO2 absorbers optimum design: from local scale to large industrial scale. CFD modelling of packed columns may be used for the calculation of total pressure drop and for the modelling of different mechanisms resulting in pressure drop. This may be used for predicting performance and for optimizing operation conditions. The information gained can also be used for improving the packing. CFD is obviously suitable for simulating flow distribution and calculating pressure drop in auxiliary column equipment like liquid and gas distributors. There seems to be no attempts in the literature to simulate an overall model for an absorption process with CFD. A major challenge is how to model the gas/liquid interface.

526

L.E. i

7. Combination of models and calculation tools 7.1. User-defined components in process simulation tools In process simulation programs, there are normally possibilities to include user-defined components. In Pro/II, it is possible to write routines in Fortran that can be linked to Pro/ II. User-defined subroutines in Fortran can also be linked to Aspen Plus. In Aspen HYSYS, new routines can be written in C++ or in Visual Basic. And in Aspen HYSYS, it is possible to introduce Excel-like spreadsheets which can both import and export data to other parts of the flowsheet calculation. An example where spreadsheets are used in Aspen HYSYS is given in Section 7.4. 7.2. Process simulation programs and Cape-Open

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

Process simulation programs are specialized in combining many types of components like equilibrium models and calculation tools for columns and other unit operations. The objective of the projects Cape-Open from 1997 to 1999 and Global Cape-Open from 1999 to 2001 was to define standard interfaces between the major components of a process simulation program. The contributors to these projects were mainly vendors of major process simulator programs and major chemical companies. One result of these programs was a set of interface standards between process modelling components and programs. The organization Cape-Open Laboratory Network (http://www.colan.org) is now maintaining the Cape-Open interface standards. 7.3. Cost estimation of CO2 removal from flue gas with amines

Most cost estimation work is performed within commercial companies and is not published. Mariz [66] has given some background for cost estimation of an Econamine process (MEA-based process from Fluor Daniel) for CO2 removal from flue gas. In the CO2 Capture Project (CCP), Choi et al. [67] performed a study with title: CO2 Removal from Power Plant Flue Gas Cost Efficient Design and Integration Study. The process simulation programs like Aspen Plus, Pro/II and Aspen HYSYS have dimensioning and cost estimation tools that can cost estimate the equipment after the process calculations. In several student projects at Telemark University College, Aspen HYSYS has been used for process simulation of a CO2 removal process followed by mechanical equipment dimensioning, cost estimation of the installed equipment and estimation of energy cost as the most important operating cost. This has normally been done by utilizing spreadsheets. Parameter variation makes it possible to find the most economical temperatures, circulation rates and column heights [68]. 7.4. Example of combination of models and calculation tools for economical optimization using HYSYS A cost estimation of the base case in Figure 2 has been performed [69], using the Aspen HYSYS version 7.0. The base case specifications from Table 1 were used, except for the Murphree efficiency, which was set to 0.15 for 1 m of packing height. Equipment cost was calculated using an open available cost estimator available on the internet by Peters et al. [70]. Installed cost (for all the equipment) was calculated using a type- and cost-dependent installation factor [69]. These dimensioning and cost estimation calculations were performed

Mathematical and Computer Modelling of Dynamical Systems

527

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

in spreadsheets connected to the process simulation program. The installed cost of the equipment was estimated to 160 mill. USD, and the energy cost for 10 years was estimated to 370 mill. USD. The absolute value of the total estimate was not regarded as very accurate, but it was expected to include most of the factors varying with size and capacity. Different parameter values were then varied to find the optimum process parameters [69]. Equipment cost change from base case to new conditions was calculated by multiplying with the capacity ratio raised to 0.65. Optimum parameters were found by performing several simulations, and optimum was found at minimum net present value (installed cost plus energy cost for 10 years with full operation). Figure 7 shows the net present (negative) value as a function of minimum temperature difference with the optimum temperature difference at 19 C. This value has high uncertainty, especially due to the uncertainty in the cost connected to the heat exchanger. Other calculated optimums were 16 m of packing height (each with Murphree efficiency 0.15) and 35 C in absorber inlet gas temperature. For the inlet temperature optimization, the Murphree efficiency was specified to be temperature dependent, varying linearly from 0.14 to 0.16 between 30 C and 50 C. These calculations were performed with much trial and error, especially due to divergence problems in the column calculations. The optimization calculation using the Optimizer tool in Aspen HYSYS [69] was also tried. A variable (here the temperature difference) was automatically varied in repeated simulations to achieve the minimum criteria function (here the net present value). For the case of minimum temperature difference, it was calculated to 17.7 C, close to the value of 19 C found in Figure 7 by specifying the parameter in each simulation. The figure shows a flat minimum between 16 C and 21 C. However, the Optimizer tool in Aspen HYSYS has some limitations, especially in optimizing column parameters. The column height cannot be optimized by Optimizer, because the number of column stages is an integer value. And the absorber inlet temperature cannot be optimized using the Optimizer tool, because the Murphree efficiencies in the

590 580 570 560 550 540 530 520 5 10 15 Delta T (C) 20 25

Figure 7. Net present (negative) value of CO2 removal plant as a function of minimum T in amine/ amine heat exchanger (from i et al. [69]).

Net present value (million USD)

528

L.E. i

column cannot be specified as temperature-dependent variables. It is an obvious challenge to make such automatic optimizations possible. There are many challenges in making robust models combining complex equilibrium models, kinetic models, column solving and optimization tools. 7.5. When is the use of rigorous models doubtful?

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

There is in general a trade-off between complex and accurate models compared to simpler and more robust models. The purpose of the calculation will of course influence on what is most important. Calculation time is not a very important factor any longer. A more important problem with a too rigorous model is when the model leads to divergence. Another drawback with a too complicated model is that it can lead to confusion and make the model more difficult to specify. An example of an unnecessary complicated model is the use of differential equations (as in Equations (12) and (13)) to calculate the absorption rate when the absorption is in the pseudo-first-order regime and can be calculated by Equation (11). Trying to solve the differential equations does not increase the accuracy and increases computing time and divergence tendency. When column calculations are in a flowsheet network, the columns are calculated several times. In process simulation programs for steady-state calculations, the column solver methods, for example the methods mentioned in Section 2.1, traditionally consist of only algebraic equations. The columns are then modelled as stages representing each plate in a plate column or a specified packing height in a packed column. More rigorous models involving differential equations describing concentration and temperature profiles in both axial and radial directions are also possible. However, a rigorous column model combined with a complex description of the equilibrium and kinetics is doubtful if this leads to flowsheet divergence. In the case of time-dependent simulation, the robustness of the models is probably more important than the accuracy. It is important to find a good balance in each flowsheet to obtain a fast, robust and accurate calculation. For direct use in mass transfer calculations, rigorous fluid flow models like CFD are probably too rigorous. Mass transfer models normally need parameters like mass transfer numbers and effective gas/liquid interfacial area, and these are traditionally not calculated in CFD calculations. More specific correlations for mass transfer numbers and interfacial area are normally more relevant. But for the future, it is of course a challenge to combine CFD with mass transfer calculations.

7.6. Validation of the flowsheet calculations containing different models and calculation tools A validation of the process flowsheet calculations against reality is difficult because there are no plants in operation at large scale (1 mill. ton CO2/yr). Data from smaller plants are not easily available. An article for validation of simulation codes against pilot plants for CO2 removal was presented in 2009 by Luo et al. [71]. Most programs (with different equilibrium models) show reasonable results compared to pilot plant data. Aspen Plus gives slightly higher heat of desorption results compared to other programs and pilot plant data. Validation of flowsheet calculations can be done by comparing with other calculations. The flowsheet calculation in Section 2.1 gives 85% CO2 removal with an energy consumption of 3.67 MJ/kg CO2 at base case conditions. When changing equilibrium model from

Mathematical and Computer Modelling of Dynamical Systems

529

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

Kent Eisenberg to Li/Mather in the Aspen HYSYS flowsheet calculation the CO2 removal was reduced to 82% and the heat consumption was reduced to 3.4 MJ/kg CO2 [11]. There are literature values for similar processes between 3.9 and 4.9 MJ/kg CO2 from Desideri and Paolucci [3] and Alie et al. [4]. Most of these energy consumption values are calculated using Aspen Plus with the Chen/Austgen model. In Liu et al. [25], it is claimed that the original electrolytic NRTL model of Chen/Austgen tends to overpredict the heat consumption and suggests new parameter values to correct this. Validation of cost estimation calculations is difficult because there are no exact cost values. The cost estimates calculated in Section 7.4 are not claimed to be very accurate. The investment cost estimate is, however, expected to include most of the cost factors that vary with size and capacity. Because of this, the calculated optimum process values like 16 m packing height, 35 C inlet temperature and 18 C in heat exchanger temperature difference are assumed to be reasonable. There are no calculated optimum values in the literature to compare with. There are, however, typical values available in the literature. In the CO2 Capture Project [67], a total absorber height of 28 m was used, where the packing height is only a part. This compares well with our calculated optimum of 16 m packing height. In the same project, the gas to the absorber was cooled to 47 C compared to our calculated optimum of 35 C. There is a high uncertainty in the optimum value of this temperature. The gas cooling cost is very dependent on local conditions. In the calculations, there is a high uncertainty in the temperature dependency of the absorption efficiency. The CO2 Capture Project uses a minimum heat exchanger temperature difference of 11 C. This is lower than our calculated optimum at 18 C. Tobiesen et al. [7] have shown in process simulations that there is little energy to save by reducing this temperature difference. This indicates that the optimum temperature difference is higher than 11 C. A main aim with this work is to show the possibility to calculate such optimum process values. 8. Summary

The tasks of modelling the CO2 removal process can be divided into descriptions of absorption and reaction kinetics, gas/liquid equilibrium, gas and liquid flows and pressure drop. Process simulation tools containing models for most of these tasks are commercially available, and the calculated results can be used as a basis for equipment dimensioning and economical optimization. A flowsheet calculation in the program Aspen HYSYS is used as an example. Calculation convergence is important, especially the column convergence is critical. For some simplified conditions, calculation of stage efficiencies can give a satisfactory description of the absorption process. There is still a challenge to search for improved vapour/liquid equilibrium models. There is need for improved accuracy, and the models should be robust and easy to use in combination with kinetic models. There is serious deviation between different estimation methods for mass transfer in structured packing. The deviation is especially large for the different prediction methods for effective gas/liquid interfacial area. CFD is an efficient tool for calculating flow conditions, pressure drop and temperature profiles, especially for one fluid phase. CFD is obviously suitable for simulating flow distribution and for calculating pressure drop in auxiliary equipment like liquid and gas distributors. It is a challenge to make more use of CFD for gas/liquid processes. An unsolved problem is the description of the gas/liquid interfacial area, and especially combined with absorption.

530

L.E. i

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

There is a traditional trade-off between complex and accurate models compared to simpler and more robust models. Under some conditions, a non-rigorous model is accurately enough. There is a challenge to find out under which conditions a simplified method is satisfactory. In time-consuming calculations like optimization and time-dependent simulations, it can be advantageous to avoid complex models that may lead to divergence, for example complex equilibrium models and models involving the solving of differential equations. In such cases a more rigorous model can be used in a separate calculation. A major challenge is the combination of different models and calculation tools. In the case of process simulation programs, Cape-Open is an example of a standard interface for introducing a new component into an existing program package. An improved model for a specific task must be available and possible to combine with other calculation tools to be utilized by other programs. In an example, models for equilibrium and mass transfer efficiency are used in a flowsheet calculation including CO2 absorption and desorption, followed by economical optimization. The example illustrates some possibilities, limitations and challenges. References
[1] A. Kohl and R. Nielsen, Gas Purification, 5th ed., Gulf Publications, Houston, 1997. [2] P.V. Danckwerts and M.M. Sharma, The absorption of carbon dioxide into solutions of alkalis and amines (with some notes on hydrogen sulphide and carbonyl sulphide), The Chemical Engineer, October (1966), pp. 244280. [3] U. Desideri and A. Paolucci, Performance modelling of a carbon dioxide removal system for power plants, Energy Conversion Manage. 40 (1999), pp. 18991915. [4] C. Alie, P. Backham, E. Croiset, and P.L. Douglas, Simulation of CO2 capture using MEA scrubbing: a flowsheet decomposition method, Energy Conversion Manage. 46 (2005), pp. 475487. [5] C. Chen and L.B. Evans, A local composition model for the excess Gibbs energy of aqueous electrolyte systems, AICHE J. 32 (1986), pp. 444454. [6] D.M. Austgen, G.T. Rochelle, X. Peng, and C. Chen, Model of vaporliquid equilibria for aqueous acid gas-alkanolamine systems using the electrolyte-NRTL equation, Ind. Eng. Chem. Res. 28 (1989), pp. 10601073. [7] F.A. Tobiesen, H.F. Svendsen, and K.A. Hoff, Desorber energy consumption amine based absorption plants, Int. J. Green Energy 2 (2005), pp. 201215. [8] A. Aroonwilas, A. Chakma, P. Tontiwachwuthikul, and A. Veawab, Mathematical modelling of mass-transfer and hydrodynamics in CO2 absorbers packed with structured packings, Chem. Eng. Sci. 58 (2003), pp. 40374053. [9] H.M. Kvamsdal, J.P. Jacobsen, and K.A. Hoff, Dynamic modelling and simulation of a CO2 absorber column for post-combustion CO2 capture, Chem. Eng. Process: Process Intensification 48 (2009), pp. 135144. [10] T. Greer, S.A. Jayarathna, M. Alic, M.C. Melaaen, and B. Lie, Dynamic model for removal of carbon dioxide from a post combustion process with monoethanolamine, 6th Vienna Conference on Mathematical Modelling (MATHMOD 09), Vienna, Austria, 2009. [11] L.E. i, Aspen HYSYS simulation of CO2 removal by amine absorption from a gas based power plant, The 48th Scandinavian Conference on Simulation and Modeling (SIMS 2007), Gteborg, Sweden, 2007. [12] R.L. Kent and B. Eisenberg, Better data for amine treating, Hydrocarbon Process. 76 (1976), pp. 8790. [13] M.H. Li and K.P. Shen, Calculation of equilibrium solubility of carbon dioxide in aqueous mixtures of monoethanolamine with methyldiethanolamine, Fluid Phase Equilib. 85 (1993), pp. 129140. [14] N.A. Al-Baghli, S.A. Pruess, V.F. Yesavage, and M.S. Selim, A rate-based model for the design of gas absorbers for the removal of CO2 and H2S using aqueous solutions of MEA and DEA, Fluid Phase Equilib. 185 (2001), pp. 3143.

Mathematical and Computer Modelling of Dynamical Systems

531

[15] S. Freguia and G.T. Rochelle, Modelling of CO2 capture by aqueous monoethanolamine, AICHE J. 49 (2003), pp. 16761686. [16] L. Kucka, I. Mller, E.Y. Kenig, and A. Gorak, On the modelling and simulation of sour gas absorption by aqueous amine solutions, Chem. Eng. Sci. 58 (2003), pp. 35713578. [17] G.F. Versteeg, L.A.J. Van Dijck, and W.P.M. Van Swaaij, On the kinetics between CO2 and alkanolamines both in aqueous and non-aqueous solutions: an overview, Chem. Eng. Comm. 144 (1996), pp. 113158. [18] M. Caplow, Kinetics of carbamate formation and breakdown, J. Am. Chem. Soc. 90(24) (1968), pp. 67956803. [19] P.M.M. Blauwhoff, G.F. Versteeg, and P.M. Van Swaaij, A study on the reaction between CO2 and alkanolamines in aqueous solutions, Chem. Eng. Sci. 39 (1984), pp. 207225. [20] E.B. Rinker, H.A. Ashour, and O.C. Sandall, Kinetics and modelling of carbon dioxide absorption into aqueous solutions of N-methyldiethanolamine, Chem. Eng. Sci. 50 (1995), pp. 755768. [21] G. Sartori and D.W. Savage, Sterically hindered amines for CO2 removal from gases, Ind. Eng. Chem. Fund. 22 (1983), pp. 239249. [22] T. Mimura, S. Shimojo, T. Suda, M. Iijma, and S. Mitsuoka, Research and development on energy saving technology for flue gas carbon dioxide recovery and steam system in power plant, Energy Conversion Manage. 36 (1995), pp. 397400. [23] T. Chakravarty, U.K. Phukan, and R.H. Weiland, Reaction of acid gases with mixtures of amines, Chem. Eng. Prog. 81 (1985), pp. 3236. [24] D. Peng and D.B. Robinson, A new two-constant equation of state, Ind. Eng. Chem. Fundam. 15(1) (1976), pp. 5964. [25] Y. Liu, L. Zhang, and S. Watanasiri, Representing vaporliquid equilibrium for an aqueous MEACO2 system using the electrolyte nonrandom-two-liquid model, Ind. Eng. Chem. Res. 38 (1999), pp. 20802090. [26] Y. Li and A.E. Mather, Correlation and prediction of the solubility of carbon dioxide in a mixed alkanol solution, Ind. Eng. Chem. Res. 33 (1994), pp. 20062015. [27] L. Kaewsichan, O. Al-Bofersen, V.F. Yesavage, and M. Sami Selim, Predictions of the solubility of acid gases in monoethanolamine (MEA) and methyldiethanolamine (MDEA) solutions using the electrolyte-UNIQUAC model, Fluid Phase Equilib. 183184 (2001), pp. 159171. [28] L. Faramarzi, G.M. Kontogorgis, K. Thomsen, and E.H. Stenby, Extended UNIQUAC model for thermodynamic modeling of CO2 absorption in aqueous alkanolamine solutions, Fluid Phase Equilib. 282 (2009), pp. 121132. [29] A. Gil-Villegas, A. Galindo, P.J. Whitehead, S.J. Mills, G. Jackson, and A.N. Burgess, Statistical associating fluid theory for chain molecules with attractive potentials of variable range, J. Chem. Phys. 106 (1997), pp. 41684186. [30] N. Mac Dowell, A. Galindo, C.S. Adjiman, and G. Jackson, Modelling the phase behaviour of the CO2+H2O+amine mixtures using transferable parameters with SAFT-VR, Presentation at AIChE Annual Meeting, Nashville, Tennessee, 813 November 2009. [31] W.K. Lewis and W.G. Whitman, Principles of gas absorption, Ind. Eng. Chem. 16 (1924), pp. 12151220. [32] R. Higbie, The rate of absorption of a pure gas into a still liquid during short periods of exposure, Trans. Am. Inst. Chem. Eng. 31 (1935), pp. 365383. [33] P.V. Danckwerts, Significance of liquid-film coefficients in gas absorption, Ind. Eng. Chem. 43 (1951), pp. 14601467. [34] W. Kays, M. Crawford, and B. Weigand, Convective heat and mass transfer, 4th ed., McGrawHill, Boston, 2005. [35] D.W. Van Krevelen and P.J. Hoftijzer, Kinetics of gasliquid reactions. Part 1: general theory, Rec. Trav. Chim. 67 (1948), pp. 563568. [36] R.A. Tomkej, F.D. Otto, H.A. Rangwala, and B.R. Morrell, Tray design for selective absorption, Gas Conditioning Conference, Norman, Oklahoma, 1987. [37] W.J. DeCoursey, Absorption with chemical reaction: development of a new relation for the Danckwerts model, Chem. Eng. Sci. 29 (1974), pp. 18671872. [38] W.J. DeCoursey, Enhancement factors for gas absorption with reversible reaction, Chem. Eng. Sci. 37 (1982), pp. 14831489. [39] F.A. Tobiesen and H.F. Svendsen, Experimental validation of a rigorous absorber model for CO2 postcombustion capture, AICHE J. 53(4) (2007), pp. 846865.

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

532

L.E. i

[40] G. Astarita, D.W. Savage, and J.M. Longo, Promotion of CO2 mass transfer in carbonate solutions, Chem. Eng. Sci. 36 (1981), pp. 581588. [41] D.P. Hagewiesche, H.A. Ashour, H.A. Al-Ghawas, and O.C. Sandall, Absorption of carbon dioxide into aqueous blends of monoethanolamine and N-methyldiethanolamine, Chem. Eng. Sci. 50 (1995), pp. 10711079. [42] L. De Leye and G.F. Froment, Rigorous simulation and design of columns for gas absorption and chemical reaction I, Comput. Chem. Eng. 10(3) (1986), pp. 493504. [43] T.K. Sherwood, G.H. Shipley, and F.A.L. Holloway, Flooding velocities in packed columns, Ind. Eng. Chem. 30, No. 7 (1938), pp. 765769. [44] J.S. Eckert, No mystery in packed-bed design, Oil Gas J. 68(34) (1970), pp. 5864. [45] K. Onda, H. Takeuchi, and Y. Okumoto, Mass transfer coefficients between gas and liquid phases in packed columns, J. Chem. Eng. Jpn. 1 (1968), pp. 5662. [46] J. Bravo and J. Fair, Generalized correlation for mass transfer in packed distillation columns, Ind. Eng. Chem. Proc. Des. Dev. 21 (1982), pp. 162170. [47] J.A. Rocha, J.L. Bravo, and J.R. Fair, Distillation columns containing structured packings: a comprehensive model for their performance. 1. Hydraulic models, Ind. Eng. Chem. Res. 32 (1993), pp. 641651. [48] J.A. Rocha, J.L. Bravo, and J.R. Fair, Distillation columns containing structured packings: a comprehensive model for their performance. 2. Mass-transfer model, Ind. Eng. Chem. Res. 35 (1996), pp. 16601667. [49] R. Billet and M. Schultes, Prediction of mass transfer columns with dumped and arranged packings, Trans. IChemE. 77 (1999), pp. 498504. [50] M.H. De Brito, U. von Stockar, and P. Bomio, Predicting the liquid phase mass transfer coefficient kL-for the Sulzer structured packing Mellapak, ICHEME Symp. Ser. 128 (1992), pp. 137144. [51] E. Brunazzi, A. Paglianti, and L. Petarca, Design of absorption columns equipped with structured packings, La Chimica e lIndustria 78 (1996), pp. 459467. [52] P. Valluri, O.K. Matar, A. Mendes, and G.F. Hewitt, Modelling hydrodynamics and mass transfer in structured packings A review, Multiphase Sci. Tech. 14(4) (2002), pp. 303348. [53] G.Q. Wang, X.G. Yuan, and K.T. Yu, Review of mass-transfer correlations for packed columns, Ind. Eng. Chem. Res. 44 (2005), pp. 87158729. [54] S. Shetti and R. Cerro, Fundamental liquid flow correlations for the computation of design parameters for ordered packings, Ind. Eng. Chem. Res. 36 (1997), pp. 771783. [55] P.A. Nawrocki, Z.P. Xu, and K.T. Chuang, Mass transfer in structured corrugated packing, Can. J. Chem. Eng. 69 (1991), pp. 13361343. [56] Z. Olujic, Development of a complete simulation model for prediction the hydraulic and separation performance of distillation columns equipped with structured packing, Chem. Biochem. Eng. Q. 11, No. 1 (1997), pp. 3146. [57] Z. Olujic, A.B. Kamerbeek, and J. de Graauw, A corrugation geometry based model for efficiency of structured distillation packing, Chem. Eng. Process. 38 (1999), pp. 683695. [58] A. Shilkin and E.Y. Kenig, A new approach to fluid separation modelling in the columns equipped with structured packings, Chem. Eng. J. 110 (2005), pp. 87100. [59] I. Iliuta and F. Larachi, Mechanistic model for structured-packing-containing columns: irrigated pressure drop, liquid holdup, and packing fractional wetted area, Ind. Eng. Chem. Res. 40 (2001), pp. 51405146. [60] M. Klker, E.Y. Kenig, and A. Grak, On the development of new column internals for reactive separations via integration of CFD and process simulation, Catal. Today 7980 (2003), pp. 479485. [61] C.F. Petre, F. Larachi, I. Iliuta, and B.P.A. Grandjean, Pressure drop through structured packings: breakdown into the contributing mechanisms by CFD modelling, Chem. Eng. Sci. 58 (2003), pp. 163177. [62] F. Larachi, C.F. Petre, I. Iliuta, and B. Grandjean, Tailoring the pressure drop of structured packings through CFD simulations, Chem. Eng. Process. 42 (2003), pp. 535541. [63] I. Iliuta, C.F. Petre, and F. Larachi, Hydrodynamic continuum model for two-phase flow structured-packing-containing columns, Chem. Eng. Sci. 59 (2004), pp. 879888. [64] L. Raynal, C. Boyer, and J. Ballaguet, Liquid holdup and pressure drop determination in structured packing with CFD simulations, Can. J. Chem. Eng. 82 (2004), pp. 871879. [65] L. Raynal, F.B. Rayana, and A. Royon-Lebeaud, Use of CFD for CO2 absorbers optimum design: from local scale to large industrial scale, GHGT-9, Energy Procedia. 1 (2009), pp. 917924.

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

Mathematical and Computer Modelling of Dynamical Systems

533

Downloaded by [Indian Institute of Petroleum] at 01:48 20 June 2012

[66] C.L. Mariz, Carbon dioxide recovery: large scale design trends, J. Can. Petrol. Tech. 37(7) (1998), pp. 4247. [67] G.N. Choi, R. Chu, B. Degen, H. Wen, P.L. Richen, and D. Chinn, CO2 removal from power plant flue gas cost efficient design and integration study, in CO2 Capture Project, Vol. 1, D.C. Thomas, ed., Elsevier, Oxford, 2005, pp. 99116. [68] T.G. Amundsen, C.H. Arntsen, E.A. Blaker, and A.M. Morland, Aspen HYSYS simulation and cost estimation of CO2 removal, MSc Project, Telemark University College, Porsgrunn, Norway, 2007. [69] L.E. i, E.A. Blaker, and N.H. Eldrup, Cost optimization of process parameters for CO2 removal by absorption in MEA using aspen HYSYS, Poster Presentation at 5th Trondheim Conference on CO2 Capture, Transport and Storage, NTNU, Trondheim, Norway, 1617 June 2009. [70] M.S. Peters, K.D. Timmerhaus, and R.E. West, Internet Estimator. Available at http://www.mhhe. com/engcs/chemical/peters/data/ (Accessed 4 June 2008). [71] X. Luo, J.N. Knudsen, D. de Montigny, T. Sanpasertparnich, R. Idem, D. Gelowitz, R. Notz, S. Hoch, H. Hasse, E. Lemaire, P. Alix, F.A. Tobiesen, O. Juliussen, M. Kpcke, and H.F. Svendsen, Comparison and validation of simulation codes against sixteen sets of data from four different pilot plants, GHGT-9, Energy Procedia 1, 2009, pp. 12491256.

Вам также может понравиться