kinetics study of carbon dioxide absorption in aqueous solutions of 1,6-hexamethyldiamine (hmda) and...

12
Kinetics study of carbon dioxide absorption in aqueous solutions of 1,6-hexamethyldiamine (HMDA) and 1,6-hexamethyldiamine, N,N 0 di-methyl (HMDA, N,N 0 ) Prachi Singh n , W.P.M. van Swaaij, D.W.F. (Wim) Brilman Faculty of Science and Technology, University of Twente, P.O. Box 217, 7500 AE, Enschede, The Netherlands article info Article history: Received 25 January 2011 Received in revised form 6 June 2011 Accepted 7 June 2011 Available online 3 July 2011 Keywords: Amine CO 2 Kinetics Acid gas Absorption Rate constant abstract A study towards the kinetics of CO 2 in aqueous solutions of 1,6-hexamethyl diamine (HMDA) and 1,6- hexamethyl diamine, N,N 0 di-methyl (HMDA, N,N 0 ) was performed at concentrations ranging from 0.5 to 2.5 mol/L and temperatures from 283 up to 303 K. The kinetics data were determined by CO 2 absorption experiments using a stirred cell reactor with a flat interface between gas and liquid. These new CO 2 solvents were identified in earlier work for their high CO 2 capacity and limited corrosiveness. The experimental technique was validated using kinetic experiments for a 2.5 mol/L monoethanola- mine solution. In view of double amine functionality and the six carbon chain between the amine groups, attention was paid to whether the amine groups acted independently and whether or not internal cyclisation would affect the carbamate forming mechanism. The reaction order with respect to HMDA was found to vary from 1.4 to 1.8 with increasing temperature. Absorption experiments in an equimolar solution of HMDA with HCl showed that the two amine groups react independently from each other towards CO 2 . The reactivity of both diamines was more than five times larger than for monoethanolamine. The secondary diamine HMDA, N,N 0 was found to be even more reactive towards CO 2 . Additionally, the effect of CO 2 loading on the kinetics was studied for 0.5 mol/L aqueous solutions of HMDA and HMDA, N,N 0 at 293 K. Both solvents are from absorption kinetics point of view good candidates for further evaluation as solvent (-component) for CO 2 capture. & 2011 Elsevier Ltd. All rights reserved. 1. Introduction Separation of carbon dioxide from a large amount of flue gas by chemical separation is known as one of the most reliable and economical processes. The separation is, however, expensive with standard industrial technology, both in terms of capital cost (capex) and operating cost (opex). The main cost items of the process are the size of the absorber and the regenerator, and the lean/rich cross flow heat exchanger. High opex is related to the energy requirement for regeneration of the solvent circulating in the process, corresponding to 80% of the total. Other issues are associated with degradation, precipitation, corrosion, foaming, evaporation, etc. Therefore, novel solvent systems are desired to be developed to make the removal of carbon dioxide from flue gases a more cost effective process. Results obtained from previous work on development of new solvents for CO 2 absorption process with aqueous amine based solvents have revealed some new novel solvents (Singh et al., 2007, 2010). The potential solvents 1,6-hexamethyldiamine (HMDA) and 1,6-hexamethyl diamine, N,N 0 di-methyl (HMDA, N,N 0 ) (see Fig. 1) showed an almost two times higher CO 2 absorption capacity when compared to conventional solvent, e.g. monoethanolamine (MEA) at similar amine concentration. For example, 0.5 mol/L 1,6-hexam- ethyldiamine, N,N 0 di-methyl reaches up to 1.5 mol CO 2 /mol amine at 10 kPa CO 2 partial pressure (similar CO 2 partial pressure in flue gas) at 30 1C temperature and 1 atm pressure. It can be noticed that HMDA and HMDA, N,N 0 solvents were found to have fairly high boiling point (HMDA¼ 475 K and HMDA, N,N 0 ¼ 484 K) when com- pared to that of MEA (444 K). Also the vapour pressure for these two solvents was found to be comparative to MEA. Water solubility for HMDA and HMDA, N,N 0 was found to be higher than that of piper- azine. Moreover, HMDA, N,N 0 is found to be non-corrosive from its MSDS data base. Hence, 1,6-hexamethyldiamine, N,N 0 di-methyl shows great potential and deserving further investigation. In order to be able to determine the dimension of the absorber in a CO 2 capture plant, information on the reaction kinetics of CO 2 with this new solvent is required. The absorption kinetics of CO 2 in aqueous solutions of 1,6-hexamethyldiamine and 1,6-hexam- ethyldiamine, N,N 0 di-methyl were determined at temperatures ranging from 283 up to 303 K and solvent concentration ranging Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/ces Chemical Engineering Science 0009-2509/$ - see front matter & 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.ces.2011.06.008 n Corresponding author. Tel.: þ31 053 489 2141. E-mail address: [email protected] (P. Singh). Chemical Engineering Science 66 (2011) 4521–4532

Upload: prachi-singh

Post on 05-Sep-2016

213 views

Category:

Documents


2 download

TRANSCRIPT

Chemical Engineering Science 66 (2011) 4521–4532

Contents lists available at ScienceDirect

Chemical Engineering Science

0009-25

doi:10.1

n Corr

E-m

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

Kinetics study of carbon dioxide absorption in aqueous solutionsof 1,6-hexamethyldiamine (HMDA) and 1,6-hexamethyldiamine,N,N0 di-methyl (HMDA, N,N0)

Prachi Singh n, W.P.M. van Swaaij, D.W.F. (Wim) Brilman

Faculty of Science and Technology, University of Twente, P.O. Box 217, 7500 AE, Enschede, The Netherlands

a r t i c l e i n f o

Article history:

Received 25 January 2011

Received in revised form

6 June 2011

Accepted 7 June 2011Available online 3 July 2011

Keywords:

Amine

CO2

Kinetics

Acid gas

Absorption

Rate constant

09/$ - see front matter & 2011 Elsevier Ltd. A

016/j.ces.2011.06.008

esponding author. Tel.: þ31 053 489 2141.

ail address: [email protected] (P. Singh

a b s t r a c t

A study towards the kinetics of CO2 in aqueous solutions of 1,6-hexamethyl diamine (HMDA) and 1,6-

hexamethyl diamine, N,N0 di-methyl (HMDA, N,N0) was performed at concentrations ranging from

0.5 to 2.5 mol/L and temperatures from 283 up to 303 K. The kinetics data were determined by CO2

absorption experiments using a stirred cell reactor with a flat interface between gas and liquid. These

new CO2 solvents were identified in earlier work for their high CO2 capacity and limited corrosiveness.

The experimental technique was validated using kinetic experiments for a 2.5 mol/L monoethanola-

mine solution. In view of double amine functionality and the six carbon chain between the amine

groups, attention was paid to whether the amine groups acted independently and whether or not

internal cyclisation would affect the carbamate forming mechanism. The reaction order with respect to

HMDA was found to vary from 1.4 to 1.8 with increasing temperature. Absorption experiments in an

equimolar solution of HMDA with HCl showed that the two amine groups react independently from

each other towards CO2. The reactivity of both diamines was more than five times larger than for

monoethanolamine. The secondary diamine HMDA, N,N0 was found to be even more reactive towards

CO2. Additionally, the effect of CO2 loading on the kinetics was studied for 0.5 mol/L aqueous solutions

of HMDA and HMDA, N,N0 at 293 K. Both solvents are from absorption kinetics point of view good

candidates for further evaluation as solvent (-component) for CO2 capture.

& 2011 Elsevier Ltd. All rights reserved.

1. Introduction

Separation of carbon dioxide from a large amount of flue gasby chemical separation is known as one of the most reliable andeconomical processes. The separation is, however, expensive withstandard industrial technology, both in terms of capital cost(capex) and operating cost (opex). The main cost items of theprocess are the size of the absorber and the regenerator, andthe lean/rich cross flow heat exchanger. High opex is related tothe energy requirement for regeneration of the solvent circulatingin the process, corresponding to 80% of the total. Other issues areassociated with degradation, precipitation, corrosion, foaming,evaporation, etc. Therefore, novel solvent systems are desired tobe developed to make the removal of carbon dioxide from fluegases a more cost effective process.

Results obtained from previous work on development of newsolvents for CO2 absorption process with aqueous amine basedsolvents have revealed some new novel solvents (Singh et al., 2007,

ll rights reserved.

).

2010). The potential solvents 1,6-hexamethyldiamine (HMDA) and1,6-hexamethyl diamine, N,N0 di-methyl (HMDA, N,N0) (see Fig. 1)showed an almost two times higher CO2 absorption capacity whencompared to conventional solvent, e.g. monoethanolamine (MEA) atsimilar amine concentration. For example, 0.5 mol/L 1,6-hexam-ethyldiamine, N,N0 di-methyl reaches up to 1.5 mol CO2/mol amineat 10 kPa CO2 partial pressure (similar CO2 partial pressure in fluegas) at 30 1C temperature and 1 atm pressure. It can be noticed thatHMDA and HMDA, N,N0 solvents were found to have fairly highboiling point (HMDA¼475 K and HMDA, N,N0 ¼484 K) when com-pared to that of MEA (444 K). Also the vapour pressure for these twosolvents was found to be comparative to MEA. Water solubility forHMDA and HMDA, N,N0 was found to be higher than that of piper-azine. Moreover, HMDA, N,N0 is found to be non-corrosive from itsMSDS data base. Hence, 1,6-hexamethyldiamine, N,N0 di-methylshows great potential and deserving further investigation.

In order to be able to determine the dimension of the absorberin a CO2 capture plant, information on the reaction kinetics of CO2

with this new solvent is required. The absorption kinetics of CO2

in aqueous solutions of 1,6-hexamethyldiamine and 1,6-hexam-ethyldiamine, N,N0 di-methyl were determined at temperaturesranging from 283 up to 303 K and solvent concentration ranging

1,6 Hexamethylenediamine (HMDA) 1,6 Hexamethylenediamine ,N,N’di-methyl (HMDA, N,N’)

NN N

NC C

Fig. 1. Molecular structure of selected diamine based solvents.

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–45324522

from 0.5 up to 2.5 mol/L. In addition, the effect of CO2 solventloading is evaluated. In order to validate the experimentalprocedure reference experiments for the kinetics of a 2.5 mol/Laqueous monoethanolamine (MEA) solution were conducted.

2. Theory

2.1. Reaction mechanism

In aqueous environment, 1,6-hexamethyldiamine and 1,6-hex-amethyldiamine, N,N0 di-methyl can react with CO2 to form manydifferent reaction products. 1,6-hexamethyldiamine is a six carbonchain length diamine with two primary amine functionalities,whereas 1,6-hexamethyldiamine, N,N0 di-methyl is a six carbonchain length diamine with two secondary amine functionalities.

In aqueous 1,6-hexamethyldiamine and 1,6-hexamethyldia-mine, N,N0 di-methyl solution, CO2 can react with two primaryamine or secondary amine group according to the followingreactions:

R1HNR2NHR1þCO2þB2R1HNR2NR1COO�þBHþ (1)

R1HNR2NR1COO�þCO2þB2COO�R1NR2NR1COO�þBHþ (2)

CO2þOH�2HCO3� (3)

CO2þH2O2H2CO3 (4)

where R1¼CH3 for HMDA, N,N0 and R1¼H for HMDA; R2¼(CH2)6;B is any base present in solution (R1HNR2NHR1, R1HNR2NR1COO� ,þR1þH2NR2NHR1, H2O, OH�).

In principle, a diamine molecule (both for HMDA as well as forHMDA, N,N0) can serve both as carbamate former and protonacceptor (base), leading to

þH2R1NR2NHR1þCO22þH2R1NR2NR1COO� (5)

The aim of this work is to determine the overall CO2 absorp-tion kinetics and to identify if possible the most importantreaction(s) and the corresponding mechanism(s) and kineticconstant(s) between CO2 and the two amine groups present in1,6-hexamethyldiamine and 1,6-hexamethyldiamine, N,N0 di-methyl. Two well-known reaction mechanisms, discussed below,are generally used to describe the reaction between CO2 andaqueous amines and are used to derive rate equations to correlatethe kinetic rate constant(s).

2.2. Termolecular mechanism

Initially proposed by Crooks and Donnellan (1989), the mecha-nism assumes that the amine bonding to CO2 and the protontransfer take place simultaneously. The mechanism was recentlyreviewed by da Silva and Svendsen (2004). The reaction of thetwo secondary group of 1,6-hexamethyldiamine, N,N0 di-methyl

with CO2 is illustrated below:

ð6Þ

In above mechanism ‘n2’ represents two lone electrons presentat nitrogen and at base atom. For the carbamate formationreactions, with HMDA (or HMDA, N,N0) and water as the dom-inating bases, the forward reaction rate can be expressed asfollows:

RCO2¼�ðkam

3 ½R1HNR2NHR1�þkH2O3 ½H2O�Þ½CO2�½R1HNR2NHR1� ð7Þ

In this equation the reaction rate expressions for the termo-lecular reaction in which water acts as a base and the one inwhich R1HNR2NHR1 acts as a base are added.

2.3. Zwitterion mechanism

Initially proposed by Caplow (1968) and reintroduced byDanckwerts (1979) this mechanism predicts the carbamate for-mation to proceed in two steps, i.e. the reaction between CO2 andthe amine proceeds through the formation of a zwitterion (reac-tion for the secondary diamine HMDA, N,N0 illustrated in (8a))and the subsequent deprotonation of the zwitterions by a base B(secondary group see (8b)):

ð8aÞ

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–4532 4523

ð8bÞ

In above mechanism ‘n2’ represents two lone electrons presentat nitrogen and at base atom. Assuming the quasi-steady-statecondition for the zwitterions concentration and an irreversibledeprotonation step, the kinetics rate equation is given by

RCO2¼

k2½CO2�½R1HNR2HNR1�

1þ ðk�1=P

kB½B�Þ¼

½CO2�½R1HNR2HNR1�

ð1=k2Þþðk�1=k2Þ ð1=P

kB½B�Þð9Þ

where SkB[B] is the contribution of all bases present in the solution(R1NHR2NHR1, R1NHR2NR1COO� ,þR1

þHNR2NHþR1, H2O, OH�) forthe removal of the protons. As, e.g., Kumar et al. (2003) pointed out,there are two asymptotic situations for amines in aqueous solution.

In case the deprotonation of the zwitterions is very fast, ork�1=

PkB½B�51, the kinetic equation reduces to simple second

order kinetics, as found for primary alkanolamine such as MEA:

RCO2¼ k2½CO2�½R1HNR2NHR1� ð10Þ

The reversed situation of case I occurs when k�1=P

kB½B�b1.Now the kinetic rate expression reduces to

RCO2¼ k2 CO2½ � R1HNR2NHR1½ �

PkB½B�

k�1

� �ð11Þ

As the reaction order is dependent on the contribution of theindividual bases to the deprotonation of the zwitterions, thisexpression can account for a shift in reaction order in the amineconcentration with changing amine concentration. This has beenfound in the kinetic rate expression for the reaction of CO2 withmany secondary alkanolamines (Versteeg et al., 1996).

Both asymptotic options I and II cannot be excluded before-hand. The pKa value of these compounds is higher than the one forMEA, which would—based on the BrØnsted plot technique—resultin high deprotonation rates and thus could point towards beha-viour type I. On the other hand, both amine groups in HMDA, N,N0

are in fact secondary amines (like diethanolamine (DEA)) for whicha reaction order for HMDA, N,N0 in excess of one is expected. In thecase of HMDA and HMDA, N,N0 a more specific situation may exist,in case a single diamine molecule simultaneously serves bothfunctions of carbamate formation and proton acceptor.

2.4. Kinetics measurement

The reaction kinetics can be determined from the absorptionrate in the reactive solution when measuring in the ‘‘pseudo-first-order reaction regime’’, with RCO2

¼ kov½CO2�. For this, the follow-ing conditions for the Hatta number (Ha) need to be satisfied(Westerterp et al., 1984):

2oHa5E1A ð12Þ

with

Ha¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffikovDCO2

pkL

ð13Þ

and

E1A ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiDCO2 ,L

DAm,L

ffiffiffiffiffiffiffiffiffiffiffiffiffiDAm,L

DCO2 ,L

s½Amine�RT

nAmmCO2PCO2 ,G

ð14Þ

assuming the penetration model to be applicable and the reac-tions to be essentially irreversible. In the above equation thestoichiometric coefficient (nAm) for HMDA and HMDA, N,N0 is one.This value is taken because both amine groups present in thesesolvents are located far apart, and, hence, it is expected that theywill not influence each others reactivity.

As the reaction of CO2 with amine is basically reversible, theinfinite enhancement factor may become lower than suggested byEq. (14). A method to calculate the infinite enhancement factor forreversible reactions is given by Secor and Beutler (1967), and ismore recently reviewed by Hogendoorn et al. (1997). The calcula-tion of the infinite enhancement factor taking into accountreversibility requires knowledge on the equilibrium constant,which is not available for the system under consideration. In thiskinetic study, at the conditions applied, it is assumed that thereaction is essentially irreversible, i.e. the equilibrium is far on theproducts side of the equation. This is supported by the low CO2

equilibrium partial pressures (lower than for equivalent loadingsfor MEA, see Singh (2011)).

For irreversible reactions, normally the Hatta number shouldbe at least five times smaller than the infinite enhancement factorto meet the condition of a pseudo-nth order regime as indicatedby Eq. (12). For reversible reactions this margin should be evenhigher. In the pseudo-first-order regime the CO2 absorption ratecan then be described by

JCO2¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffikovDCO2 ,L

q mCO2PCO2 ,G

RTð15Þ

In the absorption experiments the overall reaction rate con-stant was determined per experiment in line with the followingrate expression:

RCO2¼ kov½CO2� ð16Þ

In situation when CO2 partial pressure is increased (i.e. resultsin decrease in infinite enhancement factor), depletion of amineconcentration will start to occur at the gas–liquid interface. This isan ‘‘intermediate regime’’ in which direct calculation of thekinetic rate constant kov from the CO2 absorption flux and thecorresponding enhancement factor using Eq. (15) is not possible.An approximate solution for this intermediate regime was givenDeCoursey (1974):

EDC ¼�Ha2

2ðE1A �1Þþ

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiHa4

4ðE1A �1Þ2þ

E1A Ha2

ðE1A �1Þþ1

vuut ð17Þ

From this Ha, and, hence, kov can be determined from theexperimental determined enhancement factor E. Eq. (17) issuitable for an irreversible second order chemical reaction andis based on the surface renewal theory proposed by Danckwerts.This ‘‘intermediate regime’’ can also be used for reversiblechemical reaction with the knowledge of equilibrium constantof the reaction due to its influence on infinite enhancement factor.This regime is considered to be less attractive to derive kineticsreliably due to these mutual interactions.

For absorption rate experiments carried out in the pseudo-first-order regime, the flux should be independent of the liquid-sidemass transfer coefficient kL and thus the stirrer speed. This isgenerally a good check on the fulfilment of pseudo-first-orderregime condition in Eq. (12). Besides this, in the pseudo-first-orderregime a linear relation between the flux and the CO2 partialpressure is expected. So, even if the exact value of the infiniteenhancement factor for the reversible reactions cannot be calcu-lated, the independency of the flux on the stirrer speed and a linearrelationship between the flux and the CO2 partial pressure indi-cates that Eq. (13) is fulfilled. As can be seen in Eq. (16), thecalculation of kov from the absorption flux requires, amongst other

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–45324524

things, knowledge of the physical solubility parameter (mCO2) for

CO2 in the reactive solutions under consideration. Physical solubi-lity parameter (mCO2

) for CO2 will be determined via solubilitymeasurements of the chemically inert N2O in the reactive solutionsand using the well-established N2O–CO2 analogy.

3. Experimental procedure

Experiments were performed in a (closed) stirred cell reactor(see Fig. 2). Carbon dioxide could be fed to the reactor from two gassupply vessel that had volumes of �325 and �85 ml or directlyfrom the gas cylinder. The pressure in the reactor, which has avolume of �720 ml, could be kept constant by a pressure controller.The stirred cell, is a double walled glass reactor with thermostat,was closed by stainless steel flanges. Two separate operating stirrerswere used to stir the gas and the liquid phase separately.

A pressure transducer was connected on the reactor formonitoring the pressure in the reactor, and second one wasconnected to the gas supply vessels for determining the pressurein these vessels.

The experimental setup allowed two operation modes:

Semi-batch operation: In this operation mode the pressure in thereactor was kept constant, and the pressure drop against timewas measured in the gas supply vessel. In the pseudo-first-order

Fig. 2. Schematic representation of

regime, the relation between the pressure and the time in the gassupply vessel can be obtained from an unstationary massbalance (Hogendoorn et al., 1995):

PCO2 ,G,Supply Vessel9t ¼ PCO2 ,G,Supply Vessel9t ¼ 0

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffikovDCO2 ,L

q mCO2PCO2 ,G,Reactor

VG,Supply Vesselt ð18Þ

Total batch operation: After admittance of the batch of gas, thereactor was closed, and the pressure drop in the reactoragainst time was measured. The relation between the pressureand the time in the reactor, when closed, is given by(Blauwhoff et al., 1984)

lnPCO2 ,G,Reactor9t ¼ lnPCO2 ,G,Reactor9t ¼ 0�

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffikovDCO2 ,L

q mCO2A

VG,Reactort

ð19Þ

When the reaction rate of CO2 with the amine based absorbentsis high (as expected with secondary amine), large Hatta numberswill be obtained. A large Hatta number has the consequence thatthe CO2 partial pressure in the reactor must be very low (and hencethe total pressure close to the water vapour pressure) to satisfy thesecond condition as stated by Eq. (12) (Ha5E1A ).

the stirred cell reactor setup.

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–4532 4525

4. Physical constants

4.1. Density, viscosity and diffusion coefficient

The density of 1,6-hexamethyldiamine and 1,6-hexamethyl-diamine, N,N0 di-methyl solution with concentration ranging from1 up to 5 mol/L were measured with a commercial density metre(DMA-4500M, Anton Paar GMbH) at temperature ranging from293 up to 353 K. Table 1 shows the measured density for aqueoussolution of 1,6-hexamethyldiamine and 1,6-hexamethyldiamine,N,N0 di-methyl at different concentrations and temperatures.

The viscosities of 1,6-hexamethyldiamine and 1,6-hexamethyl-diamine, N,N0 di-methyl solution with concentration ranging from1 up to 5 mol/L was measured with an Ubbelohde-type viscometerat temperature ranging from 293 up to 353 K. The viscometer wasimmersed in a large oil bath. The temperature was controlled witha constant temperature circulator to within 70.05 K. From theefflux time, the kinematic viscosity was calculated from

n¼ Ct ð20Þ

where n is the kinematic viscosity, ‘C’ is a constant specific to theviscometer and t is the efflux time. End effect corrections wereneglected in the calculation of the kinematic viscosity. The constant

Table 1Density (kg m�3), viscosity and estimated CO2 diffusion coefficient of aqueous so

N,N0 di-methyl (HMDA, N,N0).

Solvent Concentration(mol L�1)

Temperature(K)

1,6-Hexamethylenediamine

(HMDA)

1.01 293

313

333

2.50 293

313

333

5.04 313

333

353

1,6-Hexamethylenediamine,

N,N0 di-methyl (HMDA, N,N0)

2.50 293

313

333

Table 2Liquid-side mass transfer coefficient, kL, the physical solubility parameter, mN2 O, for

1,6-hexamethylenediamine (HMDA) and 1,6-hexamethylenediamine, N,N0 di-methyl (H

Solvent Concentration(mol L�1)

Temperature (

1,6-Hexamethylenediamine

(HMDA)

0.51 298

303

313

1.01 298

303

313

2.56 298

303

313

5.04 303

313

1,6-Hexamethylenediamine,

N,N0 di-methyl (HMDA, N,N0)

2.50 298

304

315

334

‘C’ in Eq. (20) is given for each viscometer. Table 1 shows the mea-sured viscosity for aqueous solution of 1,6-hexamethyldiamine and1,6-hexamethyldiamine, N,N0 di-methyl solution at different concen-trations and temperatures. Each reported measurement was repeatedwith a maximum deviation in the kinematic viscosity of approxi-mately 70.05%.

A modified Stokes–Einstein relation given in Eq. (21) (vanHolst et al., 2009) was used to estimate the diffusion coefficient ofCO2. This CO2 diffusion coefficient will be used in the determina-tion of the kinetics using Eq. (18) or (19):

ðDCO2 ,LZ0:74ÞH2O ¼ ðDCO2 ,LZ0:74ÞAmine ð21Þ

The diffusivity of CO2 in water, as required in Eq. (22), wasdetermined with the following correlation by (Versteeg and vanSwaaij, 1988)

DCO2 ,Water ¼ 2:35� 10�6exp�2119

T

� �ð22Þ

The dynamic viscosity shown in Table 1, for 1,6-hexamethyl-diamine and 1,6-hexamethyldiamine, N,N0 di-methyl was determinedby Eq. (20). Table 1 shows the estimated diffusion coefficient for CO2

in aqueous solution of 1,6-hexamethyldiamine and 1,6-hexamethyl-diamine, N,N0 di-methyl at different concentrations and temperatures.

lution of 1,6-hexamethylenediamine (HMDA) and 1,6-hexamethylenediamine,

Viscosity (Pa s) Density(kg m�3)

CO2 diffusioncoefficient (m2 s�1)

1.8E�03 990.0 1.1E�09

1.1E�03 981.5 1.8E�09

7.0E�04 970.6 2.6E�09

3.6E�03 976.6 6.6E�10

1.9E�03 964.4 1.2E�09

1.2E�03 950.9 2.0E�09

5.2E�03 920.4 5.8E�10

2.7E�03 904.1 1.1E�09

1.7E�03 887.2 1.9E�09

1.0E�02 977.3 3.0E�10

4.6E�03 963.9 6.4E�10

2.5E�03 949.4 1.2E�09

N2O and the estimated CO2 solubility parameter, mCO2, in aqueous solutions of

MDA, N,N0).

K) kL

(m s�1)m,N2O Estimated CO2

distributioncoefficient, mCO2

– 0.56 0.76

– 0.50 0.69

– 0.43 0.61

1.2E�05 0.49 0.67

1.3E�05 0.43 0.60

1.9E�05 0.35 0.50

1.0E�05 0.50 0.68

1.2E�05 0.47 0.65

1.9E�05 0.42 0.59

1.1E�05 0.56 0.76

1.8E�05 0.58 0.80

9.0E�06 0.53 0.72

9.5E�06 0.48 0.66

1.4E�05 0.47 0.66

2.6E�05 0.43 0.64

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–45324526

4.2. Liquid-side mass transfer coefficient (kL) and physical solubility (m)

The liquid-side mass transfer coefficient and the physical solubi-lity of N2O in the actual aqueous amine solution used were measuredsimultaneously in a series of experiments at various amine concen-trations and temperatures. These experiments were performed in thesame experimental setup as mentioned above, using the batch modeprocedure for gas absorption. In a typical experiment first a knownvolume of degassed aqueous amine solution was transferred into thereactor. Once the solvent reached equilibrium with temperature andits vapour pressure is recorded, N2O gas was fed to the reactor from aN2O gas cylinder till a certain pressure is reached in the reactor. Thevolume of N2O gas supply cylinder is 320 ml.

It should be noticed that during the gas feeding time the gasand liquid phase stirrers were stopped in order to minimise theabsorption of N2O during this time period. Once the N2O gas isfed, the reactor is closed and the gas and liquid phase stirrerswere switched on. The pressure drop over time in the reactor wasrecorded. The physical solubility of N2O in the aqueous aminebased absorbents was measured in the reactor vessel, with avolume of 700 ml, which was kept at the desired temperature andfilled with the known volume of around 400 ml of solution bymeasuring the pressure drop due to the absorption of the N2O.

0.0E+00

5.0E-08

1.0E-07

1.5E-07

2.0E-07

2.5E-07

3.0E-07

3.5E-07

4.0E-07

40

PCO2 (mbar)

Flu

x (m

ole.

m-2

sec

-1)

0.5 Mole/L, 293K0.5 Mole/L, 303K1 Mole/L, 293K1 Mole/L, 303 K

0 5 10 15 20 25 30 35

Fig. 3. Comparison of CO2 absorption flux (mol m�2 s�1) in aqueous solution of

1,6-hexamethylenediamine (HMDA) at 0.5 and 1.0 mol/L and temperatures of 293

and 303 K from batch mode experiments.

Table 3Experimental data on absorption rate for aqueous solution of 1,6-hexamethylenediamin

concentration of 1,6-hexamethylenediamine (HMDA) and hydrochloric acid (HCl).

Solvent Concentration(mol L�1)

Temperature(K)

1,6-Hexamethylenediamine (HMDA) 0.51 283

0.51 295

0.51 304

1.01 284

1.01 293

1.01 304

2.56 283

2.56 293

2.56 303

1,6-Hexamethylenediamine, N, N0

di-methyl (HMDA, N, N0)

0.51 284

0.51 294

0.51 303

1,6-Hexamethylenediamineþhydrochloric

acid (HMDA:HCl)

1:1 293

The physical solubility of CO2 was estimated using CO2–N2Oanalogy (Laddha et al., 1981). Table 2 shows the results from1,6-hexamethylenediamine (HMDA) and 1,6-hexamethylenedia-mine, N,N0 di-methyl (HMDA, N,N0). It should be noticed that theliquid phase stirrer speed was kept constant during all experi-ments at approximately 130 rpm.

5. Results and discussion

5.1. Validation of the experiment setup

First experimental validation of the experiment setup wascarried out using operation mode 2 (see the Experimentalprocedure section) and using an aqueous solution of 2.5 mol/Lmonoethanolamine (MEA) at 298 K. CO2 absorption into anaqueous MEA solution was chosen in the present study becauseof its large collection of kinetic data available in literature.

The second order rate constant for 2.5 mol/L MEA solution at298 K was found to be 5.52�103 L/mol s, which implies that forMEA the k2 values determined in this study are in good agreementwith published data, especially those reported by Aboudheir et al.(2003). This confirms that the experimental equipment andprocedure can be used to determine CO2 reaction kinetics fairlyaccurate. From Eq. (15) it should be noticed that a plot of the fluxversus CO2 partial pressure at the interface will yield straight line.Fig. 3 shows results of absorption experiments at 293 and 303 Kwith varying CO2 partial pressure for two amine concentrations ofaqueous solutions of hexamethylenediamine (HMDA; at 0.5 and1 mol/L). Only experiments carried out in the regime where alinear relationship existed between pressure and flux were usedin the interpretation of the experiments into reaction kinetics, asthose experiments were carried out in the pseudo-first-orderregime. It was checked experimentally that the absorption fluxdid not change on variation of the stirrer speed (while maintain-ing a flat interface). CO2 loading for each experiment was verylow, maximally approximately 0.05 mol of CO2 per mole of amine.The same procedure was followed for the aqueous solutions of1,6-hexamethylenediamine, N, N0 di-methyl (HMDA, N, N0).

5.2. Kinetics of Aqueous solution of 1,6-hexamethylenediamine

(HMDA) and 1,6-hexamethylenediamine, N,N0 di-methyl (HMDA, N,N0)

The overall absorption kinetics of the CO2 absorption reactionwith aqueous 1,6-hexamethylenediamine (HMDA) was determinedat concentrations of 0.5, 1 and 2.5 mol/L and temperature ranging

e (HMDA), 1,6-hexamethylenediamine, N,N0di-methyl (HMDA, N,N0) and equimolar

PCO2

(mbar)

mCO2(kov DCO2

)0.5

(m s�1)

kov (s�1) Ha Einf

11.2 1.9E�03 4.4Eþ03 135 1.1Eþ03

11.2 2.7E�03 8.5Eþ03 274 1.3Eþ03

6.7 3.2E�03 1.3Eþ04 341 2.6Eþ03

8.4 3.1E�03 1.5Eþ04 243 3.2Eþ03

10.3 3.8E�03 2.5Eþ04 458 3.1Eþ03

10.6 4.3E�03 3.5Eþ04 544 3.8Eþ03

9.4 4.2E�03 8.2Eþ04 444 8.1Eþ03

8.9 5.7E�03 1.0Eþ05 837 9.5Eþ03

9.4 7.0E�03 1.2Eþ05 909 9.9Eþ03

8.3 3.1E�03 1.1Eþ04 213 1.5Eþ03

8.2 3.7E�03 1.6Eþ04 370 1.8Eþ03

9.1 4.1E�03 2.0Eþ04 431 2.0Eþ03

8.4 3.7E�03 2.4Eþ04 451 3.9Eþ03

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–4532 4527

from 283 up to 303 K. For the reaction of CO2 with aqueoussolution of 1,6-hexamethylenediamine, N,N0 di-methyl (HMDA,N,N0) kinetics was determined at 0.5 mol/L concentration andtemperature ranging from 283 up to 303 K. Table 3 shows the

5.0E+02

5.5E+03

1.1E+04

1.6E+04

2.1E+04

2.6E+04

3.1E+04

3.6E+04

4.1E+04

310

Temp. (K)

k ov

(sec

-1)

0.5 Mole/L HMDA1 Mole/L HMDA0.5 Mole/L HMDA, N, N'

280 285 290 295 300 305

Fig. 4. Effect of temperature on overall reaction rate kov (s�1) for 1,6-hexa-

methylenediamine (HMDA) and 1,6-hexamethylenediamine, N,N0 di-methyl

(HMDA, N,N0).

0.0E+00

2.0E+04

4.0E+04

6.0E+04

8.0E+04

1.0E+05

1.2E+05

1.4E+05

3.0

HMDA Concentration (Mole.L-1)

k ov

(sec

-1)

283 K295 K303 K

0.0 0.5 1.0 1.5 2.0 2.5

Fig. 5. Effect of 1,6-hexamethylenediamine (HMDA) concentration on overall

reaction rate kov (s�1) at different temperatures.

Table 4Results from the reinterpretation of the experimental kinetics date based on the DeCo

Solvent Concentration(mol L�1)

Temperature(K)

HaPF

1,6-Hexamethylenediamine (HMDA) 0.51 283 13

0.51 295 27

0.51 304 34

1.01 284 24

1.01 293 45

1.01 304 54

2.56 283 44

2.56 293 83

2.56 303 90

1,6-Hexamethylenediamine, N, N0

di-methyl (HMDA, N, N0)

0.51 284 21

0.51 294 37

0.51 303 43

1,6-Hexamethylenediamineþ

hydrochloric acid (HMDA:HCl)

1:1 293 45

overall kinetics determined by experiments for 1,6-hexamethyl-enediamine (HMDA) and 1,6-hexamethylenediamine, N,N0 di-methyl (HMDA, N,N0). A list of governing Hatta numbers andcorresponding (irreversible) infinite enhancement factors (accord-ing to Eq. (12)) is presented in Table 3.

Table 3 shows not only that in all cases the Hatta number islarger than two, but also that the ratio between the calculatedinstantaneous enhancement factor and the Hatta number is suchthat influence of diffusion on transport cannot be fully excluded.Hence, although the experimental checks on mass transfer limita-tions were negative (i.e. no effect of stirrer speed and lineardependence on CO2 partial pressure), it was decided to use theDeCoursey approach to determine the Hatta number from theexperimentally determined enhancement factor. The ratio ofthe diffusion coefficient of HMDA and HMDA, N,N0 to that ofCO2, as required for the calculation of infinite enhancementfactor, was estimated using the Wilke–Chang method.

Fig. 4 and Table 3 show the effect of temperature on theoverall rate constant for 1,6-hexamethylenediamine (HMDA) and1,6-hexamethylenediamine, N,N0 di-methyl (HMDA, N,N0). It isclear from these results that the temperature and the concentra-tion affect the kinetics for CO2 absorption. It can be noticed thatthe absorption kinetics for 0.5 mol/L HMDA, N,N0 are significantlyfaster than that for 0.5 mol/L HMDA. Fig. 5 shows the effect ofconcentration for the overall rate constant at different tempera-tures for 1,6-hexamethylenediamine (HMDA).

Starting from Eq. (9), derived for the zwitterion model, it wasfound by model analysis, when trying to fit the shapes of thecurves for the different temperature series, that k�1=

PkB½B�

decreases with the increase in temperature and amine concentra-tion, with a value in the range of around 1–10, hence intermediateto the extremes indicated by Eqs. (10) and (11). Furthermore, itappears that both amine and water play a role and that thetemperature dependency (activation energy) for the deprotona-tion reaction with the dominating base (kB) is significantly higher(around factor 10) than the one for the reverse reaction (k�1).

Table 4 shows that the determination of the kinetics usingEq. (17) gives an overview of the overall kinetics constant with anuncertainty of 10–33%. To evaluate further a graphical represen-tation between all experimentally observed enhancement factorsand the calculated Hatta numbers based on the DeCourseyapproximation, using the kinetics presented in Table 4, andphysical constants presented in Tables 1 and 2, is shown inFig. 6. It can be noticed from Fig. 6 that the experimentalabsorption rates presented in Table 3, and predictions from theDeCoursey equation are in good agreement with each other.

ursey approximation method.

-Oa

kov-PFO(s�1)

Ha-DCb

kov-DC(s�1)

k2-PFO (Lmol�1 s�1)

k2-DC (Lmol�1 s�1)

5 4.4Eþ03 138 4.4Eþ03 8.4Eþ03 8.7Eþ03

4 8.5Eþ03 307 1.1Eþ04 1.5Eþ04 2.1Eþ04

1 1.3Eþ04 373 1.5Eþ04 3.2Eþ04 3.0Eþ04

3 1.5Eþ04 241 1.5Eþ04 1.7Eþ04 1.5Eþ04

8 2.5Eþ04 490 2.8Eþ04 2.4Eþ04 2.8Eþ04

4 3.5Eþ04 601 4.2Eþ04 4.0Eþ04 4.2Eþ04

4 8.2Eþ04 434 7.9Eþ04 2.5Eþ04 3.1Eþ04

7 1.0Eþ05 857 1.1Eþ05 3.4Eþ04 4.1Eþ04

9 1.2Eþ05 966 1.4Eþ05 4.8Eþ04 5.4Eþ04

3 1.1Eþ04 220 1.1Eþ04 2.1Eþ04 2.2Eþ04

0 1.6Eþ04 413 2.0Eþ04 3.1Eþ04 3.9Eþ04

1 2.0Eþ04 506 2.8Eþ04 4.0Eþ04 5.6Eþ04

1 2.4Eþ04 473 2.6Eþ04 2.4Eþ04 2.6Eþ04

0

500

1000

10005000Exp. Enhancement factor [= Ha-PFO]

Ha-

DeC

ours

ey a

ppro

xim

atio

n

Fig. 6. Parity plot of experimental enhancement factor and Hatta number from

DeCoursey approximation using the irreversible infinite enhancement factor.

0.0E+00

1.0E+04

2.0E+04

3.0E+04

4.0E+04

5.0E+04

4

1000/Temp. (K-1)

k 2 (L

/ m

ole

sec)

HMDA, This work

EDA, Li 2007

MEA, Aboudheir 2003

2.8 3 3.2 3.4 3.6 3.8

Fig. 7. Comparison of second order reaction rate constants for HMDA, EDA

(Li et al., 2007) and MEA (Aboudheir et al., 2003), 2.5 mol/L.

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–45324528

When using the reaction equation derived for the termolecularmechanism, Eq. (7), a nearly similar fit of the data can beobtained. On the basis of the limited data set, no discriminationcan be made with respect to the reaction mechanism for carba-mate formation. The suggested bimolecular mechanism, whereone amine group of the molecule acts as proton acceptor and theother one as carbamate former, is not likely, (especially at thelower temperature) as this would suggest linear relationshipsbetween kov and the amine concentration, which was notobserved for the experiments at 283 and 295 K.

To be able to compare the corresponding second order rateconstant from the apparent kinetics kov resulting from theexperimental data and from the DeCoursey approximation withliterature data, the reaction order was taken one for HMDA andHMDAN,N0, in line with the majority of literature on the kineticsof CO2 with a wide variety of aqueous (alkanol)amines. Table 4shows the kinetic rate constant (k2) based on the experiment dataand the DeCoursey approximation method. It can be noticed fromTable 4 that the DeCoursey relation (based on an irreversibleenhancement factor) for HMDA shows similar reaction rate valueas calculated from the experimental data. The results obtainedwith the DeCoursey relation are considered to be the mostaccurate, since the pseudo-first-order criteria are may not befully satisfied (see Table 4). The second order rate constantobtained from the DeCoursey approximation is used in furtherdiscussion.

When assuming that the reaction rate for 1,6-hexamethylene-diamine is first order in the amine concentration and first order inCO2. Fig. 7 shows the comparative data for the second order rateconstant of 2.5 mol/L 1,6-hexamethylenediamine (HMDA) withthe literature value of 2.5 mol/L MEA from Aboudheir et al. (2003)and 2.5 mol/L thylenediamine (EDA) from Li et al. (2007). Fig. 7shows that 1,6-hexamethylenediamine shows an almost 10 timeshigher intrinsic kinetic rate constant than for MEA. It is interest-ing to notice that the intrinsic kinetic rate constant for the HMDA(6 carbon chain length diamine) is much higher than for the EDA(2 carbon chain length diamine).

The data presented in Table 3, for all HMDA concentrationsand temperatures, is used to study the mechanism of HMDA andCO2 reaction. A log–log plot was made between apparent kineticsrate kov versus amine concentration to identify the reaction orderwith respect to HMDA. From the experimental data, the apparentreaction order for 1,6-hexamethylenediamine (HMDA) was found

to be ranging from 1.4 (for 303 K series ) up to 1.8 (for the 283 Kseries) with an average of about 1.5.

Correlation of the apparent, indicative rate constants as afunction of temperature, obtained for HMDA and using the twodifferent mechanistic models (one based on the zwitterionmechanism and one on the basis of the termolecular mechanism)and one correlation based on simple, overall power-law kineticsresults in

Zwitterion mechanism:

k2 ¼ 2:0� 106exp �2500

T

� �ðfor k�1 ¼ 1:10�5 s�1Þ ð23Þ

kam3 ¼ 5:10�10exp �

120

T

� �ð24Þ

kH2O3 ¼ 4:5� 10�6exp �

4000

T

� �ð25Þ

Termolecular mechanism:

kam3 ¼ 2:00� 104exp �

4400

T

� �ð26Þ

kH2O3 ¼ 1:0exp �

2300

T

� �ð27Þ

Power-law kinetics:

k2 ¼ 2:5� 106exp �4400

T

� �½Am�1:5 ð28Þ

The average error using these correlations shows averagedeviations of around 15% with a maximum of around 30–40%.Above correlations are therefore only useful as first estimation fortemperatures in the range of 283–303 K and HMDA concentrationof 0.5–2.5 mol/L.

For the absorption of CO2 in aqueous solution of 1,6-hexam-ethylenediamine (HMDA) and 1,6-hexamethylenediamine, N,N0

di-methyl (HMDA, N,N0) no literature data is available for com-parison. Therefore in Table 5, the rate constant (DeCourseyapproximation) for 1,6-hexamethylenediamine (HMDA) and 1,6-hexamethylenediamine, N,N0 di-methyl (HMDA, N,N0) is shown

Table 5Second order rate kinetics of various amines.

Solvent k2

(L mol�1 s�1)pKa Temperature

(K)Concentration(mol/L)

Source

HMDA 4.2Eþ04 10.8 304 1 This work

HMDA,

N,N05.6Eþ04 11.1a 303 0.5 This work

EDA 1.7Eþ04 9.8 303 0.26–0.67 Li et al.

(2007)

DETA 3.5Eþ04 9.5þ 305 1 Hartono

et al. (2009)

Pz 9.0Eþ04 9.7 303 1 Derks et al.

(2006)

MEA 6.7Eþ03 9.3 303 1 Aboudheir

et al. (2003)

DEA 5.1Eþ03 8.9 303 1 Rinker et al.

(1996)

Experimental pKa values at 303 K taken from literature Perrin (1965). þpKa value

at 303 K estimated from correlation proposed by Hartono et al. (2009).

a pKa value at 293 K estimated by ACD/pKa software.

HMDA

HMDA N,N'*

EDA

DETA

MEA

DEA

Pz

0

1

2

3

4

5

12

pKa

ln k

2

8 8.5 9 9.5 10 10.5 11 11.5

Fig. 8. BrØnsted plot for primary and secondary amine based solvents at

303 KþpKa value at 303 K estimated from correlation proposed by Hartono

et al. (2009), *pKa value is at 293 K estimated by ACD/pKa software.

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–4532 4529

together with results for diethylenetriamine, DETA (Hartono et al.,2009), ethylenediamine, EDA (Li et al., 2007), piperazine, Pz(Derks et al., 2006) and monoethanolamine, MEA (Aboudheiret al., 2003). The results suggested that 1,6-hexamethylenedia-mine (HMDA) and 1,6-hexamethylenediamine, N,N0 di-methyl(HMDA, N,N0) show significantly faster kinetics than MEA, DEAand DETA, but lower than Piperazine, which has a cyclic structure.

Possibly, an intramolecular cyclic configuration adds to thereactivity of the HMDA species. The positive effect of cyclicstructures was earlier recognized by Cullinanne and Rochelle(2006). Fig. 8 shows the BrØnsted relationship HMDA, HMDAN,N0

and other primary and secondary amine based solvents. It shouldbe noticed that the experimental pKa (basicity) values presentedin Table 5 and Fig. 8 are taken from literature (Perrin 1665). InFig. 8 and Table 5, pKa value for DETA at 303 K was estimated bycorrelation given in literature (Hartono et al., 2009). Due tounavailability of experimental pKa data for solvent HMDA, N,N0,this pKa value for HMDA N,N0 was estimated by the use ofACD/pKa software (avg. error of 0.5%) at 293 K. In Fig. 7, the linearcorrelation between the logarithm of k2 (m3/mol s) and pKa fortemperatures up to 303 K for primary and secondary aqueousalkanolamine, as proposed by Versteeg et al. (1996) (Eq. (29)),

which is presented as a straight line:

lnk2 ¼ pKaþ17:60�7188

T

� �ð29Þ

In Fig. 8, piperazine is much higher compared to other primaryand secondary amines, as it is a cyclic amine, for which probably adifferent BrØnsted relationship exists (Derks et al., 2006). Simi-larly, the deviation of DETA could be due to the presence of twoprimary and one secondary amine group present in its molecularstructure. It can be noticed from Fig. 8 that for HMDA, HMDA, N,N0

and EDA are in the similar range and reasonably in line with thecorrelation by Versteeg and van Swaaij.

5.3. Kinetics of CO2 with protonated 1,6-hexamethylenediamine

(HMDA)

Considering the effect of molecular structure, it has beenmentioned by Albert and Serjeant (1984) that a carbon chainlength of more than 4 carbon atoms in between two diaminegroups diminishes the influence of these groups on each other. Thiseffect was identified in aliphatic diamine on the basis that for morethan four carbon atoms the basicity of both amine groups wasfound to be similar to each other (Albert and Serjeant, 1984). Theeffect of carbon chain length on the basicity of both amine groupswas confirmed by Singh (2011). The HMDA basicity of both aminegroups was estimated to be pKa1

¼10.92 and pKa2¼10.13 at 20 1C.

On this basis for 1,6-hexamethylenediamine (HMDA) the twoprimary amine groups at either end of a six carbon chain lengthin between, both amine groups should be equally reactive. To testthis, a CO2 absorption experiment was performed in an equimolar(1:1) solution of hydrochloric acid (1 mol/L) and 1,6-hexamethy-lenediamine (HMDA) (1 mol/L). Hydrochloric acid (HCl) was cho-sen, since it, being a strong acid, will protonate the most basicgroups in solutions irreversibly and thus converts all HMDAmolecules to HMDAHþ (protonated HMDA).

For simplicity reasons, it is assumed that the presence of thechloride ions does not influence the reaction (rate) or the masstransfer process, even though it must be noted that its concentra-tion is identical to the HMDAHþ concentration in solution. First,batch mode experiments were performed to determine the CO2

partial pressure at which the conditions for operating in thepseudo-first-order regime were satisfied. Apparent pseudo-first-order behaviour was found at CO2 partial pressures below about15 mbar. Again, experimental conditions were adjusted to keepthe maximum carbon dioxide loading low to minimise theinfluence of reversibility of the reaction. In all experiments, theloading never exceeded 0.002 mol CO2 per mole HMDAHþ .

Results from semi-batch mode kinetic experiment for (par-tially) protonated 1,6-hexamethylenediamine (HMDA) are shownin Tables 3 and 4. It is obvious from these results that the molardiamine concentration is more important than the number ofamine groups. Since the flux for the partially protonated aminesolution is similar to the one for the non-protonated solvent, itseems that both amine groups react individually/independentlyto CO2 as the second amine group shows a similar second orderrate constant, when the first one is protonated.

5.4. Effect of CO2 loading on kinetics

In a typical CO2 absorption process, where the amine solutioncycles between absorber and stripper, the amine based solvent isalways partially loaded with CO2. Hence, it is required to identifythe effect of such a CO2 loading on the absorption rate for thesenew solvents. Kinetic experiments were performed at 293 Kfor aqueous solutions of 1,6-hexamethylenediamine (HMDA)and 1,6-hexamethylenediamine, N,N0di-methyl (HMDA, N,N0) at

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

2.5

CO2 Loading (mole CO2 / mole amine)

k ov

(sec

-1)

0.5 Mole/L HMDA

0.5 Mole/L HMDA, N,N'

0.5 Mole/L HMDA (freeamine)0.5 Mole/L HMDA,N,N'(free amine)

0.00 0.50 1.00 1.50 2.00

Fig. 9. Effect of CO2 loading on absorption rate in aqueous solution of 0.5 mol/L

1,6-hexamethylenediamine (HMDA) and 0.5 mol/L 1,6-hexamethylenediamine,

N,N0di-methyl (HMDA, N,N0) at 293 K.

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–45324530

0.51 mol/L concentration for both solvents. These experimentswere performed in semi-batch mode and using 20 mbar CO2

partial pressure in most of the experiments. When the CO2

loading increased, the CO2 equilibrium partial pressure increasedsignificantly and the kinetic experiments were performed athigher than 20 mbar CO2 partial pressure.

Results from Table 6 and Fig. 9 show that for both HMDA andHMDA, N,N0 solvents the apparent overall rate constant decreasedrastically as CO2 loading (aCO2

) increases in the solvent. Fig. 9shows that the kov value is plotted both as determined from a totalamine concentration and as calculated from a free amine basisusing kov (free amine)¼kov/(1�aCO2

). Comparable trends for theeffect of loading on kinetics were recently published by Simonset al. (2010). For comparable CO2 loading the kinetics for 1,6-hexamethylenediamine, N,N0di-methyl (HMDA, N,N0) were foundin all cases to be higher when compared to 1,6-hexamethylene-diamine (HMDA) at the same CO2 loadings. However, the differ-ence in kinetic rate constant increases strongly with increasingloading, which may suggest that either the reaction orders in the(free) amine concentration are significantly different or that theHMDA, N,N0 species forms more bicarbonate product, resulting inhigher remaining free amine concentration at a certain loading.

The temperature dependence of the apparent forward secondorder kinetic rate constant for 1,6-hexamethylenediamine(HMDA) can be described by an Arrhenius type relationship forkinetic rate constants (DeCoursey approximation) for 1,6-hexam-ethylenediamine (HMDA). Due to the lack of experimental diffu-sion coefficient data and to avoid the influence from estimateddiffusion coefficients, the rate multiplied with diffusion coeffi-cient of CO2, k2DCO2, is plotted in Fig. 10. Piperazine correlation

Table 6Experimental data on the effect of CO2 loading on absorption rate for aqueous solutio

methyl (HMDA, N,N0).

Solvent Concentration(mol L�1)

Temperature(K)

1,6-hexamethylenediamine (HMDA) 0.51 294

0.51 294

0.51 294

0.51 294

0.51 294

0.51 293

0.51 293

0.51 293

0.51 293

0.51 293

0.51 293

1,6-hexamethylenediamine, N, N’ di-

methyl

(HMDA, N, N0)

0.51 294

0.51 284

0.51 294

0.51 294

0.51 294

0.51 294

0.51 294

0.51 294

0.51 294

0.51 294

0.51 294

0.51 294

0.51 294

0.51 294

0.51 293

0.51 294

0.51 294

0.51 294

0.51 294

for second order rate constant and diffusion coefficient withtemperature proposed by Derks et al. (2006) is also plotted inFig. 10 for comparison. The temperature dependence of theapparent forward second order kinetic rate constant and diffusioncoefficient for 1,6-hexamethylenediamine (HMDA) is described by

HMDA : k2DCO2¼ 1:2� 105exp �

6:5� 103

T

!ð30Þ

n of 1,6-hexamethylenediamine (HMDA) and 1,6-hexamethylenediamine, N,N0di-

PCO2

(mbar)

mCO2(kovDCO2)0.5

(m s�1)

CO2 loading(mole CO2/moleamine)

kov (s�1)

10.7 2.8E�03 0.06 1.1Eþ04

11.1 2.5E�03 0.18 9.0Eþ03

10.4 2.3E�03 0.30 6.8Eþ03

10.7 2.0E�03 0.41 5.4Eþ03

9.6 1.9E�03 0.51 4.6Eþ03

11.9 1.2E�03 0.64 1.9Eþ03

11.1 8.0E�04 0.87 7.7Eþ02

11.1 4.8E�04 0.99 2.7Eþ02

12.2 2.9E�04 1.10 9.5Eþ01

12.2 2.0E�04 1.21 4.7Eþ01

14.2 1.2E�04 1.23 1.6Eþ01

10.4 3.2E�03 0.25 1.3Eþ04

9.9 3.2E�03 0.30 1.2Eþ04

10.7 3.0E�03 0.46 1.3Eþ04

9.6 2.8E�03 0.57 1.1Eþ04

10.5 2.3E�03 0.66 7.2Eþ03

10.3 1.9E�03 0.81 4.8Eþ03

10.2 1.5E�03 0.91 3.1Eþ03

10.1 1.2E�03 1.04 1.8Eþ03

9.5 1.1E�03 1.17 1.6Eþ03

10.3 6.2E�04 1.30 4.6Eþ02

7.9 6.5E�04 1.39 5.1Eþ02

8.7 4.1E�04 1.48 2.0Eþ02

7.1 3.6E�04 1.55 1.5Eþ02

11.7 1.4E�04 1.59 2.2Eþ01

17.5 1.3E�04 1.62 1.8Eþ01

23.8 1.2E�04 1.66 1.7Eþ01

17.1 8.8E�05 1.69 9.0Eþ00

27.0 4.8E�05 1.72 2.7Eþ00

26.0 3.2E�05 1.75 1.2Eþ00

-12.5

-11.5

-10.5

-9.5

-8.5

-7.5

-6.5

3.65

1000/T ( K-1)

ln k

2* D

CO

2

HMDA Pz, Derks 2006

3.15 3.25 3.35 3.45 3.55

Fig. 10. Arrhenius plot of the second order rate constant for the reaction of CO2

with aqueous solution of 1,6-hexamethylenediamine (HMDA) and compared with

piperazine data from Derks et al. (2006).

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–4532 4531

The apparent temperature of the diffusion coefficient of CO2 inaqueous (amine) solutions is in the order of 2.1 K (see Versteeget al., 1996). Therefore, activation temperature of the reactionamounts to on an average 4.470.3 K. Hence, the activationenergy for HMDA was found to be close to that for MEA(36.6 kJ/mol, Aboudheir et al., 2003) and piperazine (35.0 kJ/mol, Cullinane and Rochelle, 2006).

6. Conclusion

Kinetics experiments were performed for aqueous solution ofconcentration ranging from 0.5 up to 2.5 mol/L for 1,6-hexam-ethylenediamine (HMDA) and 0.5 mol/L for 1,6-hexamethylene-diamine, N,N0 di-methyl (HMDA, N,N0) at temperature rangingfrom 283 up to 303 K, using a stirred cell reactor setup. Thereaction orders were found to be one for CO2, and for amine(1,6-hexamethylenediamine, HMDA) it varied with temperaturebetween 1.4 and 1.8 with an average of about 1.5 for the range oftemperatures investigated. Kinetic experiments using an equimo-lar concentration of 1,6-hexamethylenediamine (HMDA) andhydrochloric acid (HCl) showed that the reactivity of the secondamine group (with the first one being protonated) is similar to thereactivity of the first amine group in aqueous HMDA solutionswithout hydrochloric acid addition. This suggests that there is nomechanistic interaction between both the amine groups of onemolecule by an intramolecular cyclisation configuration in thetransition state during carbamate formation. Nevertheless, theCO2 absorption kinetics found are significantly higher than forMEA, which was taken as reference solvent.

Further, the effect of CO2 loading on the reactivity of theHMDA and HMDA, N,N0 solutions was studied and the observeddecrease in reactivity could be attributed to a large extent to thedecrease in the free amine concentration. From comparison withconventional solvents as MEA, it can be concluded on the basis ofabsorption kinetics that both 1,6-hexamethylenediamine (HMDA)and 1,6-hexamethylenediamine, N,N0 di-methyl (HMDA, N,N0)have potential as new solvent for CO2 capture.

Nomenclature

A gas–liquid interface area (m2)Di,j diffusivity of component i in j phase (m2/s)E enhancement factor (dimensionless)

E1A =Einf infinite enhancement factor defined by Eq. (15)(dimensionless)

EDC DeCoursey enhancement factor by Eq. (17)(dimensionless)

k2 second order rate constant (m3/mol s)k�1 zwitterion mechanism rate constant (s�1)kB zwitterion mechanism deprotonation rate constant,

(m3/mol s)k3B0 termolecular mechanism rate constant (m6/mol2 s)k3

am deprotonation kinetic rate constant (m3/mol s)kH2O

3 deprotonation kinetic rate constant (m3/mol s)kL liquid phase mass transfer coefficient (m/s)kov overall rate constant (/s)RCO2

rate of reaction of CO2 (mol/m3 s)mCO2

physical solubility parameter for CO2 (([CO2]L/[CO2]G)eq)(dimensionless)

Pi,j partial pressure of component I in phase j (kPa)R universal gas constant (8.3143) (J/mol K)T temperature (K)Z dynamic viscosity (mPa s)n stoichiometric coefficient in Eq. (15) (dimensionless)aCO2

CO2 loading (mole CO2/mole amine)r density (kg m�3)

Acknowledgement

This research is a part of the CATO programme, the Dutchnational research programme on CO2 Capture and Storage. CATOis financially supported by the Dutch Ministry of Economic Affairs(EZ) and the consortium partners (www.co2-cato.nl).

This work is carried out at Shell Global Solutions, Amsterdam.Special thanks to Dr. Frank Geuzebroek and all members of GSGTgroup for their help and support.

I would like to give thanks to Benno Knaken for building theexperiment setup. I would like to give sincere thanks to my latecolleague Jacco van Holst to be able to work on his experimentsetup for this study.

References

Aboudheir, A., Tontiwachwuthikula, P., Chakmab, A., Idema, R., 2003. Kinetics ofthe reactive absorption of carbon dioxide in high CO2-loaded, concentratedaqueous monoethanolamine solutions. Chemical Engineering Science 58,5195–5210.

Albert, A., Serjeant, E.P., 1984. The Determination of Ionization Constant, 3rd ed.ISBN 0-412-24290-7.

Blauwhoff, P.M.M., Versteeg, G.F., van Swaaij, W.P.M, 1984. A study on the reactionbetween CO2 and alkanolamine in aqueous solutions. Chemical EngineeringScience 39, 207–225.

Caplow, M., 1968. Kinetics of carbamates formation and breakdown. Journal of theAmerican Chemical Society 90, 6795–6803.

Crooks, J.E., Donnellan, J.P., 1989. Kinetics and mechanism of the reaction betweencarbon dioxide and amines in aqueous solution. Journal of Chemical Society ofPerkin Transactions II, 331–333.

Cullinane, J.T., Rochelle, G.T., 2006. Kinetics of carbon dioxide absorption intoaqueous potassium carbonate and piperazine. Industrial Engineering ChemicalResearch 45 (8), 2531–2545.

da Silva, E.F., Svendsen, H.F., 2004. Ab Initio study of the reaction of carbamateformation from CO2 and alkanolamines. Industrial Engineering ChemicalResearch 43, 3413–3418.

Danckwerts, P.V., 1979. The reaction of CO2 with ethanolamines. ChemicalEngineering Science 34, 443–445.

Derks, P.W.J., Kleingeld, T., van Aken, C., Hogendoorn, J.A., Versteeg, G.F., 2006.Kinetics of absorption of carbon dioxide in aqueous piperazine solutions.Chemical Engineering Science 61, 6837–6854.

DeCoursey, W.J., 1974. Absorption with chemical reaction: development of anew relation for the Dankwerts model. Chemical Engineering Science 37,1483–1489.

Hartono, A., da Silva, E.F., Svendsen, H.F., 2009. Kinetics of carbon dioxideabsorption in aqueous solution of diethyletriamine (DETA). Chemical Engi-neering Science 64, 3205–3213.

P. Singh et al. / Chemical Engineering Science 66 (2011) 4521–45324532

Hogendoorn, J.A., van Swaaij, W.P.M., Versteeg, G.F., 1995. The absorption ofcarbon monoxide in cosorb solutions-absorption rate and capacity. ChemicalEngineering Journal and the Biochemical Engineering Journal 49, 243–252.

Hogendoorn, J.A., Vas Bhat, R.D., Kuipers, J.A.M., van Swaaij, W.P.M., Versteeg, G.F.,1997. Approximation for the enhancement factor applicable factor applicableto reversible reactions of finite rate in chemically loaded solutions. ChemicalEngineering Science 52, 4547–4559.

Kumar, P.S., Hogendoorn, J.A., Versteeg, G.F., 2003. Kinetics of the reaction of CO2

with aqueous potassium salt of taurine and glycine. American Institute ofChemical Engineers (AIChE) Journal 49, 203–213.

Laddha, S.S., Diaz, J.M., Danckwerts, P.V., 1981. The N2O analogy: the solubilities ofCO2 and N2O in aqueous solutions of organic compounds. Chemical Engineer-ing Science 36, 229–230.

Li, J., Henni, A., Tontiwachwuthikul, P., 2007. Reaction kinetics of CO2 in aqueousethylenediamine, ethyl ethanolamine, and diethyl monoethanolamine solu-tions in the temperature range of 298–313 K, using the stopped-flow techni-que. Industrial Engineering Chemical Research 46, 4426–4434.

Perrin, D.D., 1965. Dissociation constants for organic bases in aqueous solution.Butterworths, London (and supplement in 1972).

Rinker, E.B., Ashour, S.S., Sandall, O.C., 1996. Kinetics and modeling of carbondioxide absorption into aqueous solutions of diethanolamine. Industrial andEngineering Chemistry Research 35, 1107.

Secor, R.M., Beutler, R.L., 1967. Penetration theory for diffusion accompanied by areversible chemical reaction with generalized kinetics. American Institute ofChemical Engineers (AIChE) Journal 13, 365–373.

Simons, K., Brilman, D.W.F., Mengers, H., Nijmeijer, K., Wessling, M., 2010. Kineticsof CO2 absorption in aqueous sarcosine salt solutions Influence of concentra-tion, temperature and CO2 loading. Industrial Engineering Chemical Research

49 (20), 9693–9702.Singh, P., Niederer, J.P.M., Versteeg, G.F., 2007. Structure and activity relationships

for amine based CO2 absorbents—I. International Journal of Greenhouse GasControl 1 (1), 5–10.

Singh, P., Brilman, D.W.F., Groeneveld, M.J., 2010. Evaluation of CO2 solubility inpotential aqueous amine-based solvents at low CO2 partial pressure. Interna-tional Journal of Greenhouse Gas Control 5 (1), 61–68.

Singh, P., 2011. Ph.D. Thesis. University of Twente, The Netherlands.van Holst, J., Versteeg, G.F., Brilman, D.W.F., Hogendoorn, J.A., 2009. Kinetic study

of CO2 with various amine acid salts in aqueous solution. Chemical Engineer-ing Science 64, 59–68.

Versteeg, G.F., van Swaaij, W.P.M., 1988. Solubility and diffusivity of acid gases(CO2, N2O) in aqueous alkanolamine solutions. Journal of Chemical and

Engineering Data 33, 29–34.Versteeg, G.F., van Dijck, L.A.J., van Swaaij, W.P.M., 1996. On the kinetics between

CO2 and alkanolamines both in aqueous and non-aqueous solutions: an

overview. Chemical Engineering Community 144, 113–158.Westerterp, K.R., van Swaaij, W.P.M., Beenackers, A.A.C.M., 1984. Chemical Reactor

Design and Operation. Wiley, New York.