You are on page 1of 122

EQUILIBRIUM AND MASS TRANSFER BEHAVIOUR OF CO2 ADSORPTION

ON ZEOLITES, CARBON MOLECULAR SIEVE, AND ACTIVATED CARBONS

A Thesis

Submitted to the Faculty of Graduate Studies and Research

In Partial Fulfillment of the Requirements for the

Degree of Master of Applied Science

In Process Systems Engineering

University of Regina

Md Ariful Islam Sarker

Regina, Saskatchewan

October, 2012

Copyright© 2012: Sarker


UNIVERSITY OF REGINA

FACULTY OF GRADUATE STUDIES AND RESEARCH

SUPERVISORY AND EXAMINING COMMITTEE

Md Ariful Islam Sarker, candidate for the degree of Master of Applied Science in Process
Systems Engineering, has presented a thesis titled, Equilibrium and Mass Transfer
Behaviour of CO2 Adsorption on Zeolites, Carbon Molecular Sieve, and Activated
Carbons, in an oral examination held on September 6, 2012. The following committee
members have found the thesis acceptable in form and content, and that the candidate
demonstrated satisfactory knowledge of the subject material.

External Examiner: Dr. Fanhua Zeng, Petroleum Systems Engineering

Supervisor: Dr. Adisorn Aroonwilas, Industrial Systems Engineering

Committee Member: Dr. David deMontigny, Industrial Systems Engineering

Committee Member: Dr. Amornvadee Veawab, Environmental Systems


Engineering

Chair of Defense: Dr. Shaun Fallat, Department of Mathematics & Statistics

*Not present at defense


ABSTRACT

Natural gas is an important source of energy that usually requires purification steps

to remove contaminants prior to pipeline transmission and industrial usage. By pressure

swing adsorption process (PSA), carbon dioxide (CO2) can be separated from natural gas

using solid materials commonly known as adsorbents. Adsorption capacity (or

equilibrium) and adsorption kinetics of the adsorption materials have great impacts on the

efficiency of CO2 removal in this PSA process. The objective of this study was to

characterize the CO2 adsorption equilibrium and kinetics of commercial adsorbents that

have potential for use in the PSA process and also to provide a better understanding of

CO2 adsorption behaviour under wide range of operating conditions.

A comprehensive set of data and analysis for CO2 adsorption equilibrium and

kinetics is presented in this study for zeolite 13x, zeolite 5A, zeolite 4A, carbon

molecular sieve (MSC-3R), activated carbon (GCA-830), and activated carbon (GCA-

1240). By using volumetric measurement technique, adsorption equilibrium and kinetic

data were taken at a temperature range of 293 – 333 K and pressure up to 35 atm. The

obtained experimental data were correlated as a function of temperature and pressure to

fit with different model equations (i.e., Langmuir, Toth, Sips, and Prausnitz). The

isosteric heat of CO2 adsorption was also estimated for individual adsorbents according to

the Clausius-Clapeyron equation. The CO2 adsorption kinetic, presented in terms of mass

transfer coefficients (k), were experimentally measured at a temperature range of 293 –

333 K and pressure up to 11 atm. The mass transfer was analyzed from the plots of CO2

uptake rate using the well-recognized linear driving force (LDF) model. The mass

i
transfer coefficients were correlated by non-linear regression to reveal the effects of

adsorption temperature and pressure. Activation energies of CO2 adsorption on the

individual adsorbents were also calculated and correlated according to the Arrhenius

equation.

ii
ACKNOWLEDGEMENTS

First of all, I would like to give my gratitude to my supervisor, Dr. Adisorn

Aroonwilas for his great support, direction, and encouragement to pursue my research

work and study. I am also thankful to Dr. Amornvadee Veawab for her significant

direction regarding my research work throughout my M.A.Sc. program.

I owe much to SaskEnergy Incorporated in Regina for their financial support. I am

grateful to the Faculty of Graduate Studies and Research at the University of Regina for

the financial support through graduate scholarships and research awards. I am also

thankful to the Faculty of Engineering and Applied Science for providing me this

opportunity.

I am also thankful to all my friends and research team at the Energy Technology

Laboratory at the University of Regina for their assistance.

Finally I would like to give thanks to my beloved parents, sisters, and brothers for

their love and support throughout my M.A.Sc. program.

iii
TABLE OF CONTENTS

Page

ABSTRACT i

ACKNOWLEDGEMENTS iii

TABLE OF CONTENTS iv

LIST OF TABLES vii

LIST OF FIGURES ix

NOMENCLATURE xii

1. INTRODUCTION 1

1.1 Need for CO2 removal from natural gas 4

1.2 Technology for CO2 separation 6

1.3 Gas adsorption technology 8

1.4 Research motivation 9

1.5 Objective and scope 11

2. CO2 ADSORPTION CHARACTERISTICS AND LITERATURE 12

REVIEWS

2.1 Adsorption equilibrium 12

2.1.1 Langmuir isotherm 13

2.1.2 Volmer Isotherm 14

2.1.3 Hill de-Boer & Fowler-Guggenheim Isotherm 14

2.1.4 Freundlich isotherm 15

2.1.5 Sips isotherm 15

iv
2.1.6 Toth isotherm 16

2.1.7 Prausnitz isotherm 16

2.1.8 Unilan isotherm 16

2.2 Adsorption kinetics 18

2.3 Isosteric heat of adsorption 22

2.4 Activation energy 23

2.5 Literature review on adsorbents for CO2 removal 24

2.5.1 CO2 adsorption by zeolite 13X 24

2.5.2 CO2 adsorption by zeolite 5A 25

2.5.3 CO2 adsorption by zeolite 4A 26

2.5.4 CO2 adsorption by carbon molecular sieve 26

2.5.5 CO2 adsorption by activated carbon 26

3. CO2 ADSORPTION EXPERIMENTS AND PROCEDURES 33

3.1 Materials 33

3.2 Experimental apparatus 34

3.3 Experimental procedures 39

3.3.1 Preparation of adsorption cell 39

3.3.2 Determination of void volume 39

3.3.3 Determination of CO2 adsorption performance 41

3.4 Experimental condition 42

3.5 Validation of experimental method 44

4. RESULTS AND DISCUSSION 46

v
4.1 CO2 adsorption equilibrium 46

4.1.1 Isotherm curves 46

4.1.2 Isotherm correlations 56

4.1.3 Isosteric heat of adsorption 64

4.2 CO2 adsorption kinetics 69

4.2.1 Mass transfer coefficient for CO2 adsorption 76

4.2.2 Mass transfer coefficient and activation energy 82

5. CONCLUSION AND FUTURE WORK 88

5.1 Conclusions 88

5.2 Recommendations for future work 90

REFERENCES 91

APPENDIX A: Experimental results of pure CO2 adsorption equilibrium 101

vi
LIST OF TABLES

Page

Table 1.1 Typical natural gas composition (in mol%) worldwide 3

Table 1.2 Typical pipeline gas specifications 5

Table 2.1 Previous works for the CO2 adsorption on zeolite 13X 28

Table 2.2 Previous works for the CO2 adsorption on zeolite 5A 29

Table 2.3 Previous works for the CO2 adsorption on zeolite 4A 30

Table 2.4 Previous works for the CO2 adsorption on carbon molecular sieve 31

Table 2.5 Previous works for the CO2 adsorption on activated carbon 32

Table 3.1 Physical properties of different adsorbents 36

Table 3.2 CO2 adsorption capacity test condition of different adsorbents 43

Table 4.1 Regression parameters for different model equations at 293 K 58

Table 4.2 Regression parameters for different model equations at 303 K 59

Table 4.3 Regression parameters for different model equations at 313 K 60

Table 4.4 Regression parameters for different model equations at 323 K 61

Table 4.5 Regression parameters for different model equations at 333 K 62

Table 4.6 Isotherm Correlation equations for adsorbents tested 63

Table 4.7 Isosteric heat of CO2 adsorption on the adsorbents 66

Table 4.8 Mass transfer coefficients of CO2 adsorption on the tested adsorbents 81

Table 4.9 Activation energy (Ea) and frequency factor (A) for the tested 85

adsorbents

Table A.1 CO2 adsorption equilibrium data on zeolite 13X at different pressure 101

vii
and temperature

Table A.2 CO2 adsorption equilibrium data on zeolite 5A at different pressure 102

and temperature

Table A.3 CO2 adsorption equilibrium data on zeolite 4A at different pressure 103

and temperature

Table A.4 CO2 adsorption equilibrium data on MSC-3R at different pressure 104

and temperature

Table A.5 CO2 adsorption equilibrium data on GCA-830 at different pressure 105

and temperature

Table A.6 CO2 adsorption equilibrium data on GCA-1240 at different pressure 106

and temperature

viii
LIST OF FIGURES

Page

Figure 1.1 Primary sources of energy in the world in 2003 2

Figure 1.2 Carbon dioxide gas removal technologies 7

Figure 2.1 Internal view of adsorbent particles 20

Figure 3.1 Photographs of adsorbents used in this study 37

Figure 3.2 Experimental apparatus for the measurement of CO2 adsorption on 38

adsorbents

Figure 3.3 Comparison of CO2 adsorption isotherm data with published data 45

for (a) zeolite 13X and (b) GCA -1240

Figure 4.1 Isotherm curves of CO2 adsorption on zeolite 13X at different 47

temperatures

Figure 4.2 Isotherm curves of CO2 adsorption on zeolite 5A at different 48

temperatures

Figure 4.3 Isotherm curves of CO2 adsorption on zeolite 4A at different 49

temperatures

Figure 4.4 Isotherm curves of CO2 adsorption on MSC-3R at different 50

temperatures

Figure 4.5 Isotherm curves of CO2 adsorption on GCA-830 at different 51

temperatures

Figure 4.6 Isotherm curves of CO2 adsorption on GCA-1240 at different 52

ix
temperatures

Figure 4.7 Comparison of isotherm curves among the adsorbents tested (a) 55

293 K; (b) 333 K.

Figure 4.8 Plots of lnP versus 1/T for (a) zeolite 13x, (b) zeolite 5A, (c) 65

zeolite 4A, (d) molecular sieve carbon MSC-3R, (e) activated

carbon GCA- 830 and (f) activated carbon GCA-1240

Figure 4.9 Correlation plot of isosteric heat of adsorption versus adsorbate 68

loading

Figure 4.10 Plots of CO2 uptake for zeolite 13X at (a) 3.4 atm; (b) 5.4 atm; (c) 70

6.8 atm; (d) 8.8 atm; (e) 10.9 atm.

Figure 4.11 Plots of CO2 uptake for zeolite 5A at (a) 3.4 atm; (b) 5.4 atm; (c) 71

6.8 atm; (d) 8.8 atm; (e) 10.9 atm.

Figure 4.12 Plots of CO2 uptake for zeolite 4A at (a) 3.4 atm; (b) 5.4 atm; (c) 72

6.8 atm; (d) 8.8 atm.

Figure 4.13 Plots of CO2 uptake for MSC-3R at (a) 3.4 atm; (b) 5.4 atm; (c) 6.8 73

atm; (d) 8.8 atm; (e) 10.9 atm.

Figure 4.14 Plots of CO2 uptake for GCA-830 at (a) 3.4 atm; (b) 5.4 atm; (c) 74

6.8 atm; (d) 8.8 atm; (e) 10.9 atm.

Figure 4.15 Plots of CO2 uptake for GCA-1240 at (a) 3.4 atm; (b) 5.4 atm; (c) 75

6.8 atm; (d) 8.8 atm; (e) 10.9 atm.

Figure 4.16 Linear plot of ln (1- ∗ ) versus time of CO2 adsorption at 293 K and 78

different pressures

x
Figure 4.17 Linear plot of ln (1- ∗ ) versus time of CO2 adsorption at 313 K and 79

different pressures

Figure 4.18 Linear plot of ln (1- ∗ ) versus time of CO2 adsorption at 333 K and 80

different pressures

Figure 4.19 Linear plot of lnk versus ( ) for CO2 adsorption activation energy 84

on (a) zeolite 13X; (b) zeolite 5A; (c) zeolite 4A; (d) MSC-3R; (e)

GCA- 830; (f) GCA-1240

Figure 4.20 Effect of pressure on activation energy and frequency factor for (a) 86

zeolite 13X; (b) zeolite 5A; (c) zeolite 4A; (d) MSC-3R; (e) GCA-

830; (f) GCA-1240

Figure 4.21 Comparison of activation energy of six different adsorbents 87

xi
NOMENCLATURE

A frequency factor or collision factor, sec-1

Å angstrom

Ap adsorption potential, kJ/mol

B affinity of the adsorbed molecules to the solid surface, atm-1

B∞ affinity constant at infinite temperature

b average affinity of the adsorbed molecules, atm-1

°C celsius

C i amount adsorbed by the component i, mol/kg

Cμ adsorbed amount of pure component i at the hypothetical

pressure, mol/kg

C T total amount adsorbed, mol/kg

CO2 carbon dioxide

Dc intra-crystalline concentration dependant diffusion

coefficient, m2/sec

DP macro-pore diffusion coefficient, m2/sec

DEA diethanol amine

DGA diglycol amine

DIPA di-isopropanol amine

Ea activation energy of adsorption, kJ/mol

∆H isosteric heat of adsorption, kJ/ mol

xii
He helium gas

k overall mass transfer coefficient, sec-1

kf external film mass transfer coefficient, m/sec

KH Henry’s constant, mol/kg.atm

k temperature dependant isotherm constant

MDEA methyl diethanol amine

MEA monoethanol amine

n total number of moles per unit mass of the adsorbent

n number of mole of pure component j per unit mass of the

adsorbent, mol/Kg

N loading of gas molecule, mol/Kg

P equilibrium pressure, atm

PS pressure sensor

PSA pressure swing adsorption

P hypothetical pressure of the pure component i, atm

P adsorption pressure of the pure component j, atm

q amount of adsorbed gas, mol/kg

q* equilibrium amount of adsorbed gas, mol/kg

∗ maximum capacity at corresponding temperature, mol/kg

R gas constant, m3 atm K−1 mol−1

RTD resistance temperature detector

RP macro particle radius, m

xiii
rc micro-crystal radius, m

s quantity of heterogeneity of the system

t temperature dependant isotherm constant

T adsorption temperature, K

TS temperature sensor

TSA thermal swing adsorption

VSA vacuum swing adsorption

w interaction energy between adsorbed molecules, kJ/mol

xi molar fraction of component i in the adsorbed phase

xj molar fraction of component j in the adsorbed phase

yi molar fraction of component i in the gas phase

yj molar fraction of component j in the gas phase

z reduced spreading pressure, mol/m3

Greek Letters

γi activity coefficient of the component i

p porosity of adsorbent particle

θ fractional coverage

π spreading pressure, kJ/m3

ɸ surface potential of the adsorbed phase per unit mass of the

adsorbent, kJ

ɸ surface potential of the pure component j, kJ

xiv
1. INTRODUCTION

The world’s primary energy sources can be classified into three categories: i) fossil

fuels such as oil, coal, and gas, ii) renewable sources including biomass, geothermal

energy, solar energy, hydro energy, tidal energy, and wind energy, and iii) nuclear energy

(Fridleifsson, 2003). Fossil fuels share the largest energy contribution, accounting for

more than 80 per cent (Figure 1.1). Burning fossil fuels for electricity and other forms of

energy has been the most accepted and widespread practice in every industry for decades,

and in some cases centuries. Recently, the use of fossil fuels has been recognized as a

major threat to the environment regarding the associated excessive greenhouse gas

emissions. The combustion of fossil fuels leases a large amount of carbon dioxide (CO2),

one of the greenhouse gases contributing to global warming and climate change. The

increasing world energy demand could lead to a substantial increase in CO2 emissions

from 26.6 gigatonnes per year in year 2003 to 40.4 gigatonnes per year by 2030

(Quadrelli and Peterson, 2007).

One strategy to reduce greenhouse gas emissions is to use so called “clean fuel”

such as natural gas since it generates relatively low CO2 emissions compared to other

fossil fuels (Mokhatab et al., 2006). Natural gas is commonly produced from

underground reservoirs. It is composed of methane (CH4) as the primary constituent and

heavier hydrocarbons such as ethane, propane, and butane, as well as non-hydrocarbons

such as nitrogen, hydrogen sulfide, helium, and CO2. The typical natural gas

compositions are listed in Table 1.1. Note that raw natural gas can contain significant

amounts of CO2 that must be removed prior to the delivery to customers.

1
Hydro, 2.1% Other, 1.2%

Biofuels, Nuclear,
4.7% 11.0% Coal/peat,
20.2%

Natural Gas,
Oil, 36.3% 24.5%

Figure 1.1 Primary sources of energy in the world in 2010

(Redrawn from key world energy statistics, IEA, 2011)

2
Table 1.1: Typical natural gas composition (in mol%) worldwide

Gas Canada Western Southwest Bach Ho Miskar Rio Arriba Cliffside


Components (Alberta) Colorado Kansas a Field County, Field,
Field
Vietnam Tunisia New Mexico Texas
Helium 0.0 0.0 0.5 0.0 0.0 0.0 1.8

Nitrogen 3.2 26.1 14.7 0.2 16.9 0.7 25.6

Carbon dioxide 1.7 42.7 0.0 0.1 13.6 0.8 0.0

Hydrogen sulfide 3.3 0.0 0.0 0.0 0.1 0.0 0.0

Methane 77.1 30.0 72.9 70.9 63.9 96.9 65.8

Ethane 6.6 0.6 6.3 13.4 3.3 1.3 3.8

Propane 3.1 0.3 3.7 7.5 1.0 0.2 1.7

Butanes 2.0 0.2 1.4 4.0 0.5 0.1 0.8

Pentanes & heavier 3.0 0.3 0.6 2.6 0.6 0.0 0.5

a
Tabular mol% data is on a wet basis (1.3 mol% water)
Source: Kidnay and Parrish, 2006

3
1.1 Need for CO2 removal from natural gas

CO2 is an incombustible gas, providing no thermal energy during the combustion

process. The presence of CO2 in natural gas product, therefore, results in the reduced

heating value per unit volume of natural gas. The presence of CO2 can also present

several challenges regarding pipeline transportation of natural gas. The large amount of

CO2 in the gas product simply occupies partial volume of the pipeline, reducing the

transport efficiency per unit volume of natural gas. In addition, CO2 in the gas product

can form carbonic acid in the presence of water and then corrode pipeline material

(Mokhatab et al., 2006). To overcome the corrosion problem, special materials are

required for process equipment and pipelines, which leads to an increase in capital

investment. If corrosion does occur due to the presence of CO2, the natural gas pipeline

system (including pipe, valves, and associated equipment) must be maintained and

replaced. The replacement downtime hinders natural gas production and delivery, causing

the production costs to rise. In some cases, natural gas product is liquefied to form so

called liquid-natural-gas (LNG) to increase transport and storage efficiency. The

liquification process usually takes place at an extremely low temperature of 112 K (-161

°C). At this temperature, if CO2 remains in the natural gas product, it will form dry ice

that freezes on heat exchanger surfaces and clogs pipelines and process equipment

(Kidnay and Parrish, 2006). To avoid all these difficulties, raw natural gas must be

processed prior to commercial use to remove CO2, and its CO2 content must be kept

within the recommended pipeline specifications. Table 1.2 shows the typical pipeline gas

specifications summarized by Younger (2004). Notice that the presence of CO2 should be

maintained below 2 vol% (or mole %).

4
Table 1.2: Typical pipeline gas specifications (Younger, 2004)

Specification Trans Alberta West coast West coast Canadian


Item Canada & Southern for US use for BC use Western
3 35.4 36.3 37.3 37.3 35.4
Minimum heating value (MJ/m )

Hydrocarbon Dew point (K at 54.4 atm) 264 264 Free of Liquids Free of Liquids ‒

3 64 64 64 64 ‒
Water Content (kg/MMm )

3 1 0.25 0.25 1 1
Hydrogen Sulphide (Grains per 2.8 m )

3 ‒ 0.20 5 ‒ ‒
Mercaptans (Grains per 2.8 m )

3 20 1 20 20 10
Total Sulphur (Grains per 2.8 m )

Carbon dioxide (mol %) 2 2 1 ‒ ‒

Oxygen content (mol %) ‒ 0.40 0.20 1 ‒

Delivery Temperature (K) 320 320 ‒ ‒ 320

Delivery Pressure (atm) 61.2 61.2 varies varies 34

5
1.2 Technology for CO2 separation

Technology selection for CO2 removal depends on the type and concentration of the

impurities in the natural gas, temperature and pressure of the feed gas, required specifications

of the gas product, volume of gas to be treated, cost (capital and operating), and

environmental regulations. From the technical viewpoint, the removal of CO2 can be

achieved by a number of approaches, as shown in Figure 1.2. Such techniques include

absorption into a liquid solvent, adsorption onto a solid adsorbent, low-temperature

cryogenic distillation, permeation through a membrane, and chemical conversion (Kohl and

Nielsen, 1997). In general, gas absorption using a physical solvent and adsorption

technologies are suitable for gas streams with high CO2 partial pressure (Kidnay and Parrish,

2006). Raw natural gas is usually available in the gas reservoirs with pressures up to 5000

psig or 340 atm, which provides considerably high partial pressure of CO2 (Younger, 2004).

The absorption technology is commonly used for processing high-volume gas streams while

the adsorption onto solids is more suitable for smaller-scale applications. Removing CO2 by

adsorption helps maintain the high pressure level of the gas stream without using an

additional compression step before delivering the methane product to the customers. Also,

adsorption technology provides low regeneration cost and non-corrosive behavior, which

makes this technology more applicable for the removal of CO2 from natural gas. In this

study, focus has been limited to the adsorption process for separating CO2 from high-

pressure natural gas.

6
MEA

DGA

Amine DEA

DIPA

MDEA
Chemical
Benfield

Catacarb
Alkali
Solvent Salt Vetro-
Absorption coke
Selexol
Flexorb
HP
Rectisol
Physical
Ifpexol
Hybrid
Purisol
PSA
Solid
TSA
CO2 Removal Adsorption
Technology Cryogenic VSA
Cellulose
acetate
Poly-
Membrane imide
Poly-
sulfone
Direct
Stretford
Conversion
Molecular
Gate

Figure 1.2 Carbon dioxide gas removal technologies

(Redrawn from Kohl and Nielsen, 1997)

7
1.3 Gas adsorption technology

Gas adsorption is a mass-transfer process that at least one selective component in a gas

mixture is driven towards and retained on the surface of adsorbent particles by Van der

Waals or electrostatic forces. The efficiency of a gas adsorption system depends on several

factors including pore size of adsorbent material, partial pressure of adsorbed component,

system temperature, and interaction between adsorbed component and adsorbent material. In

addition, the efficiency of an adsorption system is also controlled by how well the adsorbent

material can be regenerated under specific conditions. Based on the regeneration mechanism,

adsorption technology can be classified into four categories: pressure swing adsorption

(PSA), thermal swing adsorption (TSA), vacuum swing adsorption (VSA), and electrical

swing adsorption (Thomas and Crittenden, 1998). In PSA and VSA processes, the

regeneration is achieved by reducing the bed pressure after adsorption service. In TSA and

electrical swing processes, the regeneration is done by heating the adsorption bed using

either thermal energy or electric current. Among these, PSA is suitable for bulk separation

and for the system having weak adsorptive force between adsorbed component and adsorbent

material. The continuous operation of a PSA process is commonly achieved through a cyclic

procedure consisting of four sequential steps: i) pressurisation by feed gas, ii) adsorption at

high pressure, iii) counter-current depressurisation, and iv) purge with product fraction

(Thomas and Crittenden, 1998).

Selecting the most suitable adsorbent materials for a PSA process depends on a number

of adsorbent attributes including selectivity, regeneration performance, inter-particle

diffusivity, adsorption capacity, packing density, physical and chemical stability, and cost.

The adsorbent materials commonly used in the gas separation industry are zeolite-based

8
adsorbents and carbon-based adsorbents. Recently, metal organic framework adsorbents are

receiving increasing attention due to their excellent pore volume, as pore volume along with

proper pore diameter increases the total available adsorption surface area which results in

increase in gas adsorption capacity (Tagliabue et al., 2009). Zeolite-based adsorbents can be

classified into different categories depending upon pore size. The common zeolite-based

adsorbents for gas separation applications are zeolite 13X, zeolite 5A, and zeolite 4A.

Carbon-based adsorbents exhibit low polarity and heat of adsorption. The typical carbon-

based adsorbents for gas separation are carbon molecular sieves and activated carbons.

1.4 Research motivation

The design and operation of the CO2 adsorption process depends on characteristics of

the adsorbents used. To select the most appropriate adsorbent for any application, it is

extremely important to analyze equilibrium data, which are the key screening criteria

revealing the maximum amount of gaseous component to be removed per unit mass of the

adsorbent. Kinetic properties are even more important for adsorbent selection because they

provide information about the rate of adsorption under controlled conditions. Both

equilibrium and kinetic information help ensure that the removal of CO2 from the natural gas

succeeds through the use of appropriate adsorbent.

At present, there are a number of adsorbents capable of removing CO2 from natural

gas. As mentioned earlier, zeolites, carbon molecular sieves, and activated carbons have great

potential to remove CO2 from natural gas by pressure swing adsorption. In this work, zeolite

13X, zeolite 5A, zeolite 4A, a carbon molecular sieve (MSC 3R), and two activated carbons

(GCA-830 and GCA-1240) were selected to test their performance for CO2 separation from

9
natural gas. These adsorbent samples are commercially available and cover a wide variety of

adsorbent families.

In recent years, a number of research projects associated with the adsorption properties

of CO2 gas on different adsorbents were reported. Most of these projects focused on

adsorption equilibrium (or adsorption isotherms) at different temperatures and pressures. The

CO2 adsorption isotherms of zeolite 13X were measured by Cavenati et al.(2004), Zhang et

al. (2010), Siriwardane et al. (2005), and Ko et al. (2003). The CO2 adsorption capacity of

zeolite 5A was reported by Pakseresht et al. (2002), Chen et al. (1990), and Kim et al. (1995).

Siriwardane et al. (2001) showed the CO2 adsorption capacity of zeolite 4A. Castello et al.

(2004) and Bae and Lee (2005) reported the capacity of carbon molecular sieves. For the

kinetics of CO2 adsorption, there is only a limited number of projects reporting data at

service pressure of natural gas. Zhang et al. (2010), Rutherford et al. (2003), and Bae and Lee

(2005) investigated the kinetics of zeolite 13x at high CO2 feed pressures. Rutherford et al.

(2003) and Bae and Lee (2005) reported the CO2 adsorption kinetics for carbon molecular

sieves under moderate pressure. On the other hand, Rutherford and Do (2000) and Yucel and

Ruthven (1980) reported the kinetics of CO2 adsorption on zeolite 5A below atmospheric

pressure. Perez and Armenta (2010) and Yucel and Ruthven (1980) presented CO2 uptake

curves for zeolite 4A below atmospheric pressure. Because both adsorption rate and

adsorption capacity are vital for evaluating the performance and practicality of the CO2

adsorption process, the behaviour of CO2-adsorption kinetics and equilibrium must be

characterized systematically for the selected adsorbents.

10
1.5 Objective and scope

The objective of this study is to characterize the CO2 adsorption equilibrium and CO2

adsorption kinetics of the six adsorption materials selected to determine the most suitable

adsorbent for CO2 separation from natural gas using the pressure swing adsorption process.

An adsorption equilibrium and kinetic study was performed on six adsorbents (i.e., zeolite

13X, zeolite 5A, zeolite 4A, carbon molecular sieve (MSC 3R), and two activated carbons

(GCA-830 and GCA-1240)). The results of this research are expected to provide better

understanding of CO2 gas adsorption using different adsorbents, and the adsorbents’ potential

in applicability for CO2 gas separation from natural gas. The following research tasks were

carried out to attain the objectives of this study:

 Adsorption isotherms and the kinetics of CO2 adsorption on the selected adsorbents

were measured at different temperatures and pressures. The measured data were

correlated with the conventional adsorption isotherm models.

 The mass transfer coefficient and adsorption activation energy of CO2 adsorption

were analyzed. Also, the isosteric heat of adsorption was determined.

This thesis consists of five chapters. The basic theories of adsorption equilibrium and

kinetics as well as literature review of CO2 adsorption, are provided in Chapter 2. Details of

the experimental set up, materials used, experimental procedures, and test conditions for the

CO2 adsorption experiments are described in Chapter 3. Chapter 4 presents the experimental

results of the adsorption equilibrium and kinetics studies, as well as the data analysis. This

chapter also describes the isosteric heat of CO2 adsorption. Conclusions and

recommendations for future work are given in Chapter 5.

11
2. CO2 ADSORPTION CHARACTERISTICS AND LITERATURE REVIEW

Carbon dioxide adsorption is a mass-transfer process taking place when a gas stream

containing CO2 and particles of a porous material (or adsorbent) are brought into direct

contact in order to allow the CO2 gas to travel towards and then reside on the surface of

adsorbent particles. This process is exothermic, occurring by means of diffusion driven by a

non-equilibrium condition or a difference in CO2 partial pressure between the gas and solid

phases. In most cases, the adsorbent material is packed in a series of adsorption columns

where the gas stream is introduced at a regular time period until the adsorbent reaches its

saturation point. The saturated column then undergoes the desorption operation during which

the adsorbed CO2 is released under controlled pressure and temperature, and the adsorbent

capacity is restored for more adsorption. The performance of CO2 adsorption depends on two

primary characteristics: adsorption equilibrium and adsorption kinetics. The following

sections provide some basic background of these two important features.

2.1 Adsorption equilibrium

Adsorption equilibrium is a very important adsorption characteristic because it reveals

the ability of solid adsorbent to accommodate the adsorbed gaseous molecules under specific

conditions. The equilibrium data are usually presented for a given gas-solid pair as a function

of temperature (T) and pressure (P).

q* = f (P, T) (2.1)

where q is the amount of gas adsorbed per gram of adsorbent. The adsorption equilibrium is

commonly reported at a given temperature as the relationship between the partial pressure of

12
adsorbed molecules in the gas phase and the adsorbed amount q. This relationship is referred

to as the adsorption isotherm, which can be classified into different types, according to the

International Union of Pure and Applied Chemistry (IUPAC). For CO2 adsorption, there are

a number of mathematical models proposed to describe experimental adsorption isotherm

data. Some models have theoretical foundations and others are empirical in nature. The key

isotherm models are highlighted below.

2.1.1 Langmuir isotherm

This is the simplest and the most recognized theoretical isotherm model. It illustrates

monolayer adsorption on an ideal surface where the surface energy fluctuates periodically

(Do, 1998). The periodic fluctuation indicates that the adsorption surface is homogeneous

and the adsorption energy is constant and distributed evenly over all adsorption sites. It is

assumed that there is no interaction between the adsorbed molecules (Yang, 2003). The

Langmuir isotherm was derived from the dynamic-equilibrium concept (i.e., both adsorption

and desorption activities take place at the same rate per unit of surface area). The Langmuir

isotherm equation can be written as (Langmuir, 1918; Morse et al., 2010):

q* = (2.2)

where B = B∞ exp ( ) (2.3)

Here, q* is the amount of gas adsorbed, P is the equilibrium pressure, T is the adsorption

temperature, qm is the maximum amount of gas adsorbed, B∞ is the affinity constant at

infinite temperature, Q is the heat of adsorption, and R is the gas constant. The parameter, B,

is the Langmuir constant measuring the affinity of the adsorbed molecules to the solid

surface. The higher the value of B, the greater the affinity. The adsorbed amount of gaseous

13
component on adsorbent surface for a binary system can be expressed by equation (2.4),

which is known as the extended Langmuir model (Markham and Benton, 1931).

= ∑
(2.4)
,

The adsorbed amount of the species "i" (qi) can be calculated using this model equation for a

multi-component system.

2.1.2 Volmer isotherm

In 1925, Volmer proposed an isotherm model based on molecular mobility on the

adsorbent surface. The Volmer isotherm equation can be written as (Ruthven, 1984):

BP = exp (2.5)

Here, θ is the fractional coverage that can be defined as a ratio of the adsorbed amount q to

the maximum amount q*, P is the equilibrium pressure, and B is the affinity of the adsorbed

molecules to the solid. Unlike the Langmuir model, the parameter B in the Volmer model

decreases with the amount of molecules adsorbed on the adsorbent (Do, 1998). This suggests

that a change in adsorption pressure has an impact on the interaction between adsorbed

molecules.

2.1.3 Hill de-Boer & Fowler-Guggenheim isotherm

Based on an equation of state describing the adsorbent surface, two isotherm models

taking into account the mobility and interaction between adsorbed molecules were proposed

by Hill (1946), De Boer (1953), and Fowler-Guggenheim (1965). These isotherm models can

be written as:

Hill de-Boer: BP = exp exp (-cθ) (2.6)

14
Fowler-Guggenheim: BP = exp (-cθ) (2.7)

where c= (2.8)

Here, z is a co-ordination number and w is the interaction energy between adsorbed

molecules.

2.1.4 Freundlich isotherm

This is the first known empirical equation that can fit the adsorption isotherm data

deviating from the ideal situation due to the complexity of the adsorbent surface. The

mathematical form of the Freundlich model is (Siriwardane et al., 2005; Freundlich, 1907):

q* = /
(2.9)

where and t are isotherm parameters depending on the adsorption temperature (t>1). The

Freundlich isotherm can be applied to adsorption systems with heterogeneous adsorbent.

This model does not follow Henry’s law behaviour at low pressure and presents no finite

limit at the higher pressure (Do, 1998).

2.1.5 Sips isotherm

This model is the combined form of the Langmuir and Freundlich isotherms. In the

Freundlich model, the amount of gas adsorbed is increased indefinitely with pressure. A

combined model is proposed by Sips (1948) to avoid this limitation. The mathematical form

of this model is:

( ) /
q* = /
(2.10)
( )

where q*, B and t are isotherm parameters. The Sips model is transformed from the empirical

Freundlich isotherm into the theoretical Langmuir isotherm predicting the ideal adsorbent

surface only when the parameter t is equal to unity.

15
2.1.6 Toth isotherm

This model is a semi-empirical isotherm with three parameters. It can be used to

describe an adsorption system with sub-monolayer coverage, and it can also predict the

adsorption behaviour of gases at both low and high pressure. The mathematical form of this

model is (Toth, 1971; Cavenati et al., 2004):

q* = /
(2.11)
( )

In 1996, the Toth model was modified by Keller et al. (1996) to include a pressure function

instead of a constant parameter.

2.1.7 Prausnitz isotherm

In 1972, Radke & Prausnitz proposed an empirical equation having three parameters to

calculate the adsorbed amount; it can be written as (Radke & Prausnitz, 1972):


= + (2.12)

where q represents the adsorbed amount, a represents Henry’s constant, B indicates affinity

constant and t indicates Freundlich constant. The Prausnitz isotherm equation reduces to

Henry’s equation at lower adsorption pressure and the Freundlich equation at higher

adsorption pressure. Equation 2.12 can be applied successfully to a wide range of adsorption

pressures and loadings.

2.1.8 Unilan isotherm

This model takes into account the topography of the adsorbent surface, demonstrating

the surface heterogeneity. Based on the topographic data, this model equation that can be

written as (Do, 1998):

q* = ln ( ) (2.13)

16
where the parameter indicates the average affinity of the adsorbed molecules and s defines

the quantity of heterogeneity of the system.

It should be noted that the above isotherm models were developed from systems with a

single gaseous component. In multi-component systems, these equations are used together

with a set of thermodynamic equations (or a thermodynamic framework) so as to calculate

total adsorbed amount as along with the adsorbed amounts of individual gases. A

thermodynamic framework known as “Ideal Adsorption Solution Theory (IAST)” was

proposed by Myers and Prausnitz (1965) in order to represent the ideal behaviour of multi-

component gas adsorption. The ideal framework can be written as (Do, 1998):

=∑ (2.14a)

= = -∫ d (2.14b)

Pyj = xj (2.14c)

∑ =1 (2.14d)

where n is the total number of moles per unit mass of the adsorbent, xj is the molar fraction of

component j in the adsorbed phase, is the number of mole of pure component j per unit

mass of the adsorbent, ɸ is the surface potential of the adsorbed phase per unit mass of the

adsorbent, ɸ is the surface potential of the pure component j , R is the gas constant, T is the

temperature, is the adsorption pressure of the pure component j, P is the equilibrium

pressure, and yj is the molar fraction of component j in the gas phase.

According to Costa et al. (1989), the ideal framework (IAST) cannot offer accurate

prediction of adsorption of hydrocarbons and CO2 on zeolite-based adsorbents. Real

Adsorption Solution Theory (RAST), which includes an activity coefficient for non-ideality

17
of systems, can provide better prediction. The equations of the RAST model can be written

as (Costa et al., 1981; Talu and Zwiebel, 1986):

Pyi = xi γi ( ) (Z) (2.15a)

∑ =1 (2.15b)

( )
z= =∫ dP0 (2.15c)

( )
=∑ +∑ X (2.15d)

Cµi = xi CµT (2.15e)

where P is the equilibrium pressure, yi is the molar fraction of component i in the gas phase,

xi is the molar fraction of component i in the adsorbed phase, γi is the activity coefficient of

the component i, z is the reduced spreading pressure, is the hypothetical pressure of the

pure component i, is the adsorption potential, π is the spreading pressure, is the

adsorbed amount of pure component i at the hypothetical pressure, C T is the total amount

adsorbed, and C i is the amount adsorbed by component i.

2.2 Adsorption kinetics

Adsorption kinetics or adsorption rate is another important mass transfer characteristic

needed for the design and operation of a gas adsorption column. An adsorbent with large

adsorption capacity (or equilibrium) might not be the best choice for industrial applications if

its adsorption activity takes place at a very slow rate. In most cases, the adsorbent offering a

fast adsorption rate is considered to be the most suitable material.

Because typical adsorbents are porous particles, the overall adsorption rate is usually

controlled by the diffusion of adsorbate molecules from the gas phase into the interior of
18
adsorption sites (Do, 1998). In general, the diffusion of these molecules involves three

sequential steps: i) transport to the external surface of the solid particles, ii) diffusion through

the macropores, and iii) diffusion into micropores. Figure 2.1 shows a simplified diagram of

typical adsorbent particles and the general structure associated with molecular diffusion.

From the figure, it is clear that the diffusion of adsorbate molecules is subject to three mass

transfer resistances: external film resistance and macropore and micropore diffusive

resistances. Under no slip conditions at a solid boundary, it can be assumed that an adsorbent

particle is surrounded by a laminar film through which mass transfer takes place by diffusion

(Ruthven, 1984). Typically, this external resistance is negligibly small compared to the other

two resistances. Macropore resistance is usually a rate determining step for the adsorption

process. Macropore diffusion depends on the pore size of the particular adsorbent and the

nature of the fluid-wall interaction.

19
Figure 2.1 Internal view of adsorbent particles

(Modified from Do, 1998).

20
The overall mass transfer resistance (1/k), which was first proposed by Glueckauf and

Coates (1947) and later modified by Haynes and Sharma (1975), can be written in a

mathematical form as (Farooq et al., 2002):

= + + (2.16)

where the first, second, and third terms on the right represent external film resistance,

macropore resistance, and micropore resistance, respectively. Here, k is the overall mass

transfer coefficient, kf is the external film mass transfer coefficient, KH is the Henry’s

constant, RP is the macroparticle radius, DP is the macropore diffusion coefficient, Dc is the

intra-crystalline concentration dependant diffusion coefficient, rc is the microcrystal radius,

and p is the porosity of adsorbent particle.

It should be noted that the determination of individual mass transfer resistances is

rather difficult in real practice. Under typical circumstances, it is more practical to express

the adsorption rate in terms of the overall coefficient k and the overall mass transfer driving

forces across both gas and solid phases. The overall mass transfer coefficient k can be

analyzed from the uptake profiles that demonstrate the amount of gas adsorbed on the

absorbent particles as a function of time. Today, there are a number of models, such as the

linear driving force model (LDF), pore diffusion model, and dual resistance model, proposed

to represent the uptake-rate behaviour. In this study, the LDF model was chosen to analyze

the mass transfer coefficient and adsorption kinetics. The LDF model was derived from the

pore diffusion model. The following equation was proposed for the LDF model (Glueckauf

and Coates, 1947; Zhang et al., 2010):


= k (q*‒ q) (2.17)

21
where is the rate of mass transfer or adsorption rate, q* is the equilibrium amount of gas

adsorbed, q is the amount of gas adsorbed with respect to any particular time, and, again, is

the overall mass transfer coefficient. It should be noted that the coefficient k is a function of

adsorption pressure and temperature. By integrating Equation (2.17), the LDF model can be

written as:


ln = ‒ kt (2.18)


The mass transfer coefficient can be determined from the slope of a plot of 1− ∗

versus adsorption time (t).

2.3 Isosteric heat of adsorption

The isosteric heat of adsorption is an important parameter that reveals the level of

energy released during adsorption activity. The heat of adsorption could have a great impact

on the adsorption rate because the released exothermic energy can cause the temperature of

the adsorbent particles to rise, reducing the adsorption capacity (Do, 1998). The isosteric heat

of adsorption (∆H) can be calculated from the following equation Clausius-Clapeyron

equation (Hill, 1949; Lee et al., 2002):


=( ) (2.19)

Here, T is the adsorption temperature, P is the pressure, N is the loading of the gas molecule,

and R is the molar gas constant. This equation was derived from the Clausius-Clapeyron

equation under the assumption that adsorbed phase volume is negligible and the gas phase is

ideal. With a series of isotherm curves obtained at different temperatures, it is possible to

22
extract the equilibrium pressure as a function of adsorption temperature at any given loading

of gas molecule (N). The heat of adsorption can be determined from the slope of a plot of

versus reciprocal of temperature (1/T).

2.4 Activation energy

The adsorption activation energy is the potential energy barrier that must be overcome

to cause adsorption activity on the solid surface. In a physical adsorption process, activation

energy decreases as pressure increases because, at a higher pressure, adsorbate-adsorbent

interaction becomes stronger due to an increase in gas density. The activation energy of

adsorption (Ea) can be expressed using the following Arrhenius equation (Zhang et al.,

2010):

/
= (2.20)

where k is the overall mass transfer coefficient, A is the frequency factor or collision factor, R

is the molar gas constant, and T is the temperature. Equation (2.20) can also be rearranged as:

lnk = - + (2.21)

The activation energy (Ea) can be analyzed directly from the slope of a plot between and

reciprocal of temperature (1/T).

23
2.5 Literature review on adsorbents for CO2 removal

This section provides a review of studies that were carried out to evaluate the

performance of potential CO2 removal adsorbents in terms of equilibrium, kinetics, and other

associated adsorption characteristics.

2.5.1 CO2 adsorption by zeolite 13X

Most research studies on CO2 adsorption by zeolite 13X have been aimed at

determining the adsorption isotherm at low pressure ranges close to atmospheric pressure (up

to 1.2 atm). Costa et al. (1991) measured the CO2 adsorption isotherm at different

temperatures and up to 1 atm. Calleja et al. (1994) also measured the isotherm at the low

pressure range (up to 0.9 atm) and fitted the obtained data with different isotherm models.

Chue et al. (1995) measured the CO2 adsorption isotherm so as to evaluate the performance

of zeolite 13X in a pressure swing adsorption process. The measurement was made up to

1.05 atm. Lee et al. (2002) investigated the CO2 adsorption equilibrium on zeolite 13x at low

pressure (1 atm). The isosteric heat of adsorption was also analyzed from the obtained

equilibrium data. Li et al. (2008) and Zhang et al. (2009) reported the use of zeolite 13X for

removing CO2 from flue gas containing humidity and other impurities. They also provided

the adsorption isotherm at 1.2 atm and breakthrough analysis of CO2 removal. Wang and

Levan (2009) measured the adsorption isotherm at different temperatures and up to 1 atm.

The obtained data were fitted with the Toth model.

There are only a few studies focusing on CO2 adsorption by zeolite 13X at medium

pressure ranges. Harlick and Tezel (2004) investigated the adsorption capacity and the heat

of adsorption for a number of adsorbents including zeolite 13X at pressures up to 2.5 atm.

Merel et al. (2008) compared the adsorption performance between zeolites 13X and 5A at

24
pressures up to 4.9 atm and reported the adsorption isotherm, as well as breakthrough curves,

of CO2 adsorption.

For studies at high pressure ranges, Cavenati et al. (2004) measured the CO2 adsorption

isotherm and analyzed the isosteric heat of adsorption at pressures up to 49.4 atm. Zhang et

al. (2010) investigated the adsorption equilibrium and kinetics at different temperatures and

up to 29.6 atm.

2.5.2 CO2 adsorption by zeolite 5A

The adsorption of CO2 by zeolite 5A has been studied since the 1970s. In 1980, Yucel

and Ruthven investigated the adsorption isotherm and kinetics at pressure as low as 0.4 atm.

The adsorption kinetics, activation energy, and heat of adsorption were reported by Triebe

and Tezel (1995) for the removal of CO2 from air. Pakseresht et al. (2002) investigated

adsorption equilibrium at different temperatures at intermediate pressure (up to 9.9 atm). Tlili

et al. (2009) reported the adsorption equilibrium and breakthrough curve at 1 atm. Saha et al.

(2010) studied the adsorption equilibrium and kinetics of CO2 adsorption from air and

methane at 1.05 atm. Finally, Liu et al. (2011) investigated the equilibrium isotherm at

different temperatures and the pressure up to 1 atm.

For the high pressure range, Chen et al. (1990) reported the adsorption isotherm at

pressures up to 54.5 atm. Kim et al. (1995) studied the isotherm, heat of adsorption, and

breakthrough curves for removal of CO2 from H2 gas at 27.5 atm.

25
2.5.3 CO2 adsorption by zeolite 4A

Eagan and Anderson (1975) investigated the CO2 adsorption isotherm for zeolite 4A at

pressures up to 1.02 atm. Yucel and Ruthven (1980) measured the CO2 adsorption kinetics at

different temperatures at 0.4 atm. Siriwardane et al. (2001) reported the use of zeolite 4A for

the separation of CO2 from a high pressure flue gas stream at 20.4 atm. Ahn et al. (2004)

investigated the equilibrium isotherm and kinetics at 0.8 atm.

2.5.4 CO2 adsorption by carbon molecular sieve

Kikkinides and Yang (1993) reported the equilibrium capacity of a carbon molecular

sieve for CO2 adsorption from flue gas. Mochida et al. (1995) investigated the CO2

adsorption isotherm of a carbon molecular sieve for the removal of CO2 from methane at 1

atm. Also, for CO2 removal from methane, the adsorption equilibrium and kinetics were

reported by Jayaraman et al. (2002) at 5.1 atm and by Rutherford et al. (2003) at 2.4 atm. The

isotherm and kinetics of CO2 removal from air were studied at different temperatures by Reid

and Thomas (1999).

For the high pressure range, Amoros et al. (1998) examined the effect of pore size of a

carbon molecular sieve on the CO2 adsorption performance at 39.5 atm. Castello et al. (2004)

reported the CO2 adsorption isotherm at 29.6 atm. Bae and Lee (2005) investigated the

adsorption equilibrium and kinetics at 14.9 atm.

2.5.5 CO2 adsorption by activated carbon

Most research studies on CO2 adsorption by activated carbon have been carried out at

high pressure ranges. Amoros et al. (1996) studied the CO2 adsorption characteristics of

26
activated carbon at pressures up to 39.5 atm. Dreisbach et al. (1999) measured and modeled

the adsorption isotherm at 59.2 atm. The use of activated carbon for removing CO2 from the

flue gas was investigated by Siriwardane et al. (2001) at 20.4 atm and by Millward and

Yaghi (2005) at 44.4 atm. They also reported the corresponding CO2 adsorption equilibrium.

Sudibandriyo et al. (2003) investigated CO2 adsorption on activated carbon at a very high

pressure of 133.3 atm. Drage et al. (2009) reported the equilibrium capacity and kinetics of

CO2 adsorption from synthetic gas at 39.5 atm.

For low pressure applications, Chue et al. (1995) investigated the CO2 adsorption

isotherm at 1.05 atm. Guo et al. (2006) measured and modeled the isotherm and isosteric heat

of adsorption at 3.9 atm. Vaart et al. (2000) reported the CO2 adsorption isotherm at a

pressure of 7.9 atm.

Details of the above literature reviews summarized according to the type of adsorbent

materials are given in Tables 2.1 through 2.5.

27
Table 2.1: Previous works on CO2 adsorption on zeolite 13X

Author Technique Experimental conditions Isotherm Mass Transfer Isosteric Heat of Activation
Temperature Pressure coefficient Heat adsorption energy
(K) (atm) (1/sec) (kJ/mol) (kJ/mol) (kJ/mol)
Costa et al., 1991 Volumetric 279, 293 & 308 0-1 X - - - -

Calleja et al., 1994 Volumetric 293 0-1 X - - - -

Chue et al., 1995 Volumetric 288 0-1 X - - - -

Lee et al., 2002 Volumetric 273-353 0-1 X - X - -

Cavenati et al., 2004 Gravimetric 298, 308 & 323 0 - 49 X - X - -

Harlick & Tezel, 2004 Micrometrics 295 0-2 X - - X -

Siriwardane et al., 2005 Volumetric 393 0 - 20 X - - X -

Merel et al., 2008 - 298 & 323 0-5 X - - - -

Li et al., 2008 Micrometrics 293, 313 & 363 0-1 X - - - -

Zhang et al., 2009 Micrometrics 293, 313 & 363 0-1 X - - - -

Wang & Levan, 2009 Volumetric 228 - 448 0-1 X - - - -

Zhang et al., 2010 Gravimetric 298 - 328 0 - 30 X X - - X

X = Data Available

28
Table 2.2: Previous works on CO2 adsorption on zeolite 5A

Author Technique Experimental conditions Isotherm Mass Transfer Isosteric Heat of Activation
Temperature Pressure coefficient Heat adsorption energy
(K) (atm) (1/sec) (kJ/mol) (kJ/mol) (kJ/mol)
Yucel & Ruthven, 1980 Gravimetric 193 - 372 0 - 0.5 X - - - X

Chen et al., 1990 Volumetric 298 0 - 54 X - X - -

Kim et al., 1995 - 288 - 313 0 - 27 X - - X -

Triebe & Tezel, 1995 Volumetric 243 - 383 - - - - X -

Pakseresht et al., 2002 Volumetric 303 - 573 0 - 10 X - - X -

Tlili et al., 2009 Gravimetric 298 - 473 0-1 X - - - -

Saha et al., 2010 - 298 0-1 X - - - -

Liu et al., 2011 Gravimetric 303 - 423 0-1 X - X - -

X = Data Available

29
Table 2.3: Previous work on CO2 adsorption on zeolite 4A

Author Technique Experimental conditions Isotherm Mass Transfer Isosteric Heat of Activation
Temperature Pressure coefficient Heat adsorption energy
(K) (atm) (1/sec) (kJ/mol) (kJ/mol) (kJ/mol)
Eagan & Anderson, 1975 Volumetric - 0-1 X - - - -

Yucel & Ruthven, 1980 Gravimetric 273 - 371 0 - 0.5 X X - - X

Siriwardane et al., 2001 Volumetric 298 0 - 20 X - - - -

Ahn et al., 2004 Gravimetric 273 - 313 0 - 0.9 X X - - X

X = Data Available

30
Table 2.4: Previous works on CO2 adsorption on carbon molecular sieves

Author Technique Experimental conditions Isotherm Mass Transfer Isosteric Heat of Activation
Temperature Pressure coefficient Heat adsorption energy
(K) (atm) (1/sec) (kJ/mol) (kJ/mol) (kJ/mol)
Mochida et al., 1995 Volumetric 303 0- 1 - - - - -

Amoros et al., 1998 Gravimetric 273 0 - 39 X - - - -

Reid & Thomas, 1999 Gravimetric 303 - 343 0- 1 X X X - -

Jayaraman et al., 2002 Gravimetric 303 - 343 1- 5 X X - X X

Rutherford et al., 2003 Volumetric 323 & 343 0- 3 X X - X X

Castello et al., 2004 Gravimetric 273 0 - 30 X - - - -

Tan & Ani, 2004 Micrometrics 298 - X - - - -

Rodil et al., 2005 Volumetric 273 - X X - - -

Castello et al., 2005 Gravimetric 298, 313 & 328 0-1 X X - X -

Bae & Lee, 2005 Volumetric 293, 303 & 313 0 - 15 X X - - -

X = Data Available

31
Table 2.5: Previous works on CO2 adsorption on activated carbon

Author Technique Experimental conditions Isotherm Mass Transfer Isosteric Heat of Activation
Temperature Pressure coefficient Heat adsorption energy
(K) (atm) (1/sec) (kJ/mol) (kJ/mol) (kJ/mol)
Chue et al., 1995 Volumetric 288 - 343 0- 1 X - - - -

Amoros et al., 1996 Gravimetric 298 0 - 39 X - - - -

Dreisbach et al., 1999 Gravimetric 298 0 - 59 X - - - -

Vaart et al., 2000 Gravimetric 292 - 349 0- 8 X - - - -

Siriwardane et al., 2001 Volumetric 298 0 - 20 X - X - -

Sudibandriyo et al., 2003 Gravimetric 318 0 - 134 X - - - -

Millward & Yaghi, 2005 Volumetric - 0 - 44 - - - - -

Guo et al., 2006 Vacuum 303 - 333 0 - 0.5 X - X - -

Drage et al., 2009 Volumetric 303 0 - 39 X - - - -

X = Data Available

32
3. CO2 ADSORPTION EXPERIMENTS AND PROCEDURES

Adsorption experiments were carried out in this study to measure the adsorption

equilibrium or isotherm and the kinetics or uptake rate of CO2 on the selected commercial

adsorbents under different temperatures and pressures. Details of materials, experimental

equipment, and experimental procedures are provided in this chapter.

3.1 Materials

In this study, six commercial adsorbents were used for CO2 adsorption experiments.

They are zeolite 13X, zeolite 5A, zeolite 4A, a carbon molecular sieve (MSC 3R), and

granular activated carbons (GCA-830 and GCA-1240). The three zeolite adsorbents were

purchased from Sigma-Aldrich Company Limited (Oakville ON, Canada). The carbon

molecular sieve MSC 3R and granular activated carbons were donated by Japan

EnviroChemicals Ltd. (Tokyo, Japan) and Norit Activate Carbon Company Ltd (Texas,

USA), respectively.

The zeolite 13X used in this study is a pellet type of which the molecular formula is

1 Na2O: 1 Al2O3: 2.8 ± 0.2 SiO2: xH2O. It is a sodium-modified molecular sieve with a

pore diameter of 10 angstroms (Å). Zeolite 5A is a calcium form of molecular sieve with

a pore diameter of 5 (Å), and its molecular formula is Ca/nNa12-2n [(AlO2)12(SiO2)12]

·xH2O. The zeolite 4A is a sodium-modified molecular sieve having a smaller pore size

compared to zeolite 5A. It is a bead type with a particle size of 8-12 mesh, and its

molecular formula is Na12 [(AlO2)12(SiO2)12] · xH2O. The carbon molecular sieve (MSC

3R) used in this study has uniform super-micropores with a diameter of less than 10 Å.

33
For activated carbons, both GCA-830 and GCA-1240 were made from coconut shells by

steam activation. They can be regenerated by thermal reactivation. The particle sizes are

8-30 mesh for GCA-830 and 12-40 mesh for GCA-1240. The physical properties of these

adsorbents were obtained from the manufacturers and are presented in Table 3.1. Their

photographs are given in Figure 3.1.

The helium and CO2 gases used in this study were ultra pure with purity grades of

99.999%. They were purchased from Praxair Company Limited.

3.2 Experimental apparatus

The adsorption isotherm and uptake rate (kinetics) for all adsorbents were measured

using a volumetric gas adsorption apparatus designed and fabricated in the Energy

Technology Laboratory, University of Regina. A schematic diagram and a photograph of

this setup are given in Figures 3.2 (a) and 3.2 (b), respectively. The adsorption apparatus

consisted of two stainless steel (SS 304) pressure vessels purchased from Swagelok. One

vessel (Part No.: 304L-HDF4-150-T-PD) was used for gas storage, supplying gaseous

components such as CO2 and He to another vessel serving as an adsorption cell. The

capacity of the storage vessel was 150 ml with an accuracy of ±5%, and the vessel could

sustain pressure up to 122.5 atm. The adsorption cell was a packed column filled with the

adsorbent of interest. Both storage and adsorption vessels were equipped with highly

accurate pressure transmitters and RTD temperature probes. The pressure transmitter

(Cole Parmer, Part No: 68072-58) connected to the storage vessel was capable of

measuring pressure up to 34 atm with an accuracy of ± 0.05 atm, and another transmitter

(Cole Parmer, Part No: 68072-60) connected to the adsorption cell could measure up to

34
68 atm with an accuracy of ± 0.1 atm. The temperature probes for both the storage vessel

and adsorption cell had an accuracy of ± 0.15 K or (±0.15°C). Both pressure transmitters

and temperature probes were connected to a data acquisition system (OMEGA, Model

No: OM-420, Serial No: 09082411) and a computer, enabling accurate measurements of

pressure and temperature over time. Such real-time measurements enabled monitoring of

the rate and amount of gaseous component being transferred between the two vessels. A

rotary vacuum pump capable of generating a 0.78 atm vacuum (Cole Parmer, Model No:

L-79200-00) was connected to the adsorption cell to allow the regeneration of the used

adsorbents and also to help evacuate any unwanted gas held by the adsorbent prior to the

adsorption experiments. A stainless steel pressure regulator (Swagelok, Model No:

KCYIJRH 425A96050) was installed between the storage vessel and adsorption cell so as

to allow the control of feed pressure set for the adsorption cell. The entire system was

assembled using Swagelok fittings and tested to be leak proof. It could be operated at up

to 393 K and 40.8 atm. The temperature of the adsorption cell was controlled with a

temperature-controlled water bath with a precision of ±0.01 K or ±0.01 °C and set point

accuracy of ±1 K or ±1 °C (TECHNE, Model No: FTE10DP, serial No: 137097-10). An

electric oven (VMR, Canada) was used for treating the adsorbents prior to the adsorption

experiments. An auto-calibrated microbalance (Ohaus Corporation, Model: EP214C) was

used for weighing the adsorbent samples.

35
Table 3.1: Physical properties of different adsorbents

Adorbent Physical properties References


Zeolite 13X Nominal pore diameter : 8 Å Sigma Aldrich
-3
Particle Diameter : 1.6 x 10 m
3
Bulk density : 640.7 kg/m
Crush Strength : 3.2 kg
5 2
Surface area : 7.2 x 10 m /kg
-4 3
pore volume : 3.2 x 10 m /kg
Zeolite 4A Pore diameter : 4Å Sigma Aldrich
3
Bulk density : 720.8 kg/m
Regeneration temp : 473-588 K
Heat of Adsorption : 4.18 MJ/kg H2 O
Zeolite 5A Pore diameter : 5Å Sigma Aldrich
3
Bulk density : 704.8 kg/m
Regeneration temp : 473-588 K
Heat of Adsorption : 4.18 MJ/kg H2 O
6 2
GCA-830 Surface area : 1.15 x 10 m /kg Norit Americas Inc.
3
Bulk density : 464.5 kg/m
6 2
GCA-1240 Surface area : 1.15 x 10 m /kg Norit Americas Inc.
3
Bulk density : 496.6 kg/m
-3
MSC-3R Pellet diameter : 1.8 X 10 m Japan Enviro-Chemicals, Ltd.

36
(a) Zeolite 13x (b) Zeolite 5A

(c) Zeolite 4A (d) MSC-3R

(e) GCA-830 (f) GCA-1240

Figure 3.1: Photographs of adsorbents used in this study (original in colour)

37
Computer
Data Acquisition
PS System

Pressure
TS Regulator PS
(Pressure Sensor)
Adsorption TS
Storage Cell
Cell (Temperature
Sensor)

CO2 He Vacuum Manual Thermostatic


Pump valve Water Bath

(a) Schematic diagram of gas adsorption apparatus

Pressure Gas
Regulator Storage

Data
Thermostatic
Acquisition
Water Bath

Vacuum
Pump
Adsorption
Cell

(b) Photograph of the experimental setup (original in colour)

Figure 3.2: Experimental apparatus for the measurement of CO2 adsorption on

adsorbents

38
3.3 Experimental procedures

The pressure difference technique was used to determine the CO2 adsorption

isotherms and uptake rate. The amount of gas adsorbed in the adsorption cell could be

evaluated from the change in pressure of the storage vessel at a fixed temperature. As

CO2 from the storage vessel was adsorbed by the adsorbent in the adsorption cell, the

pressure of the storage vessel decreased rapidly and then became stable as soon as the

adsorption cell reached pressure equilibrium. The difference between the initial pressure

and final pressure was used for isotherm analysis while the reduction in pressure with

time was used for uptake rate analysis. Details of the experimental procedures starting

from adsorbent preparation are given below.

3.3.1 Preparation of adsorption cell

Prior to the adsorption experiment, the selected adsorbent was heated in a

convection oven so as to remove any moisture and unwanted gases. After being heated

overnight, the adsorbent was cooled down at room temperature in a desiccator. The

adsorbent was then weighed and loaded into the adsorption cell. The loaded adsorption

cell was kept under vacuum to minimize the amount of gas retained within the void

volume of the bed.

3.3.2 Determination of void volume

After the adsorption cell was properly prepared, its void volume was determined

and further used for estimating the amount of gaseous component that did not

participated in the adsorption process but occupied this empty space during each

adsorption experiment. The measurement of void volume was done through the

39
displacement technique using helium gas supplied from the storage vessel. Using the data

acquisition (DAQ) system, pressure and temperature of the storage were measured and

recorded before and after the displacement. The difference in pressure and temperature

was translated into the void volume ( ) using the following ideal gas equations:

∗ ∗ ∗
Vvoid = (3.1)

= − = − (3.2)

where Pf and Tf are the pressure and temperature of the adsorption vessel, respectively; Zf

is the compressibility factor at Pf and Tf; R is the universal gas constant; is the

number of moles of helium gas released from the storage vessel at the initial and final

conditions; is the volume of the storage vessel; and P1, T1, and Z1 are the

pressure, temperature, and compressibility factor of the storage vessel at the initial

conditions while P2, T2, and Z2 are those at the final conditions. The compressibility

factor can be calculated from the following equations (Salehi et al., 2007):

Z = 1+( + 0.01 )( ) (3.3a)

.
B0 = 0.083 − . (3.3b)

.
B1 = 0.139 – . (3.3c)

Pr = (3.3d)

Tr = (3.3e)

where Tc and Pc are the critical temperature and critical pressure, respectively.

40
3.3.3 Determination of CO2 adsorption performance

To measure the adsorption performance of the adsorbent, the storage vessel was

first charged with a CO2 gas, and its pressure and temperature were recorded as the initial

conditions. The CO2 gas from the storage was then introduced into the adsorption cell

with a specific pressure controlled by the pressure regulator. At this point, adsorption

activity started to take place in the adsorption cell, as was observed from the reduction in

pressure of the storage vessel. This adsorption activity continued until the storage

pressure reached a new equilibrium value, which was set as the final pressure and

temperature. Depending upon the test conditions, it could take from minutes to hours to

reach the new pressure equilibrium. Note that the pressure reduction of the storage vessel

was measured and recorded continuously as a function of time. This real-time

information was used for calculating the CO2 adsorption uptake rate. The following are

the equations for calculating the amount of gas adsorbed on solid adsorbent ( ):

nadsorbent = − (3.4)

nvoid = (3.5)

where is the number of moles of CO2 released from the storage vessel, is the

number of moles of CO2 that occupies the void space in an adsorption vessel, Peq and Teq

are the pressure and temperature of an adsorption vessel under equilibrium or final

conditions, and Zeq is the compressibility factor of the adsorbed gas under equilibrium

conditions.

41
3.4 Experimental conditions

Experimental conditions for the measurement of adsorption isotherm and kinetics

are given in Table 3.2.

42
Table 3.2: CO2 adsorption capacity test conditions for different adsorbents

Adsorbents Adsorption Isotherm Adsorption kinetics


Temperature (K) Pressure (atm) Temperature (K) Pressure (atm)
Zeolite 13X 293, 303, 313, 0 - 35.4 293, 313 & 333 3.4, 5.4, 6.8,
323 & 333 8.8 & 11
Zeolite 5A 293, 303, 313 0 - 35.5 293, 313 & 333 3.4, 5.4, 6.8,
& 333 8.8 & 11
Zeolite 4A 293, 303, 323 0 - 35.4 293, 313 & 333 3.4, 5.4, 6.8,
& 333 & 8.8
MSC-3R 293, 303, 313, 0 - 35.4 293, 313 & 333 3.4, 5.4, 6.8,
323 & 333 8.8 & 11
GCA-830 293, 303, 313, 0 - 35.6 293, 313 & 333 3.4, 5.4, 6.8,
323 & 333 8.8 & 11
GCA-1240 293, 303, 313, 0 - 35.4 293, 313 & 333 3.4, 5.4, 6.8,
323 & 333 8.8 & 11

43
3.5 Validation of experimental method

The experimental method was validated by comparing the CO2 adsorption

isotherms obtained in this study with those reported in the open literature. Based on the

availability of the published data, the comparison was made at 293 K and pressure up to

34 atm for two adsorbents (i.e. zeolite 13X and activated carbon GCA-1240). It is clear

from Figure 3.3 that the CO2 adsorption isotherms in this study remained within the range

of the published data, thereby validating the experimental method and analysis carried

out in this study. Figure 3.3 also indicates significant deviations in the isotherm data

among the published works as the difference in adsorbent physical property (such as pore

volume, pore diameter and available surface area) results in the variation of CO2

adsorption capacity on the adsorbent surface.

44
8.0

Amount of adsorbed CO2


6.0
(mmol/g)

4.0
This Study
Cavenati et al., 2004
2.0 Ko et al., 2005
Siriwardane et al., 2005
Zhang et al., 2010
0.0
0 10 20 30 40
Pressure (atm)
(a)

10
Amount of adsorbed CO2

8
(mmol/g)

This Study
4
Berlier & Frere, 1997
Dreisbach et al., 1999
2 Himeno et al., 2005
Goetz et al., 2006
0
0 10 20 30 40
Pressure (atm)
(b)

Figure 3.3: Comparison of CO2 adsorption isotherm data with published data for (a)

zeolite 13X and (b) GCA -1240

45
4. RESULTS AND DISCUSSION

This chapter presents the experimental results of CO2 adsorption on the selected

adsorbent materials: zeolite 13X, zeolite 5A, zeolite 4A, carbon molecular sieve (MSC-

3R) and two activated carbons (GCA-830 and GCA-1240). The CO2 adsorption

experiments were conducted under test conditions provided earlier in Chapter 3. The

results reported here can be categorized into two main aspects: i) CO2 adsorption

equilibrium indicating the maximum capacity of each adsorbent material and ii) CO2

adsorption kinetics indicating the rate of adsorption activity. The effects of process

parameters (temperature and pressure) on the adsorption behaviour are also discussed in

this chapter. The equilibrium data were analyzed further for heat of adsorption while the

kinetic data were analyzed for mass transfer coefficient and activation energy.

4.1 CO2 adsorption equilibrium

4.1.1 Isotherm curves

In this study, the CO2 adsorption equilibrium for each adsorbent was measured

under a series of isothermal conditions (i.e., 293 K, 303 K, 313 K, 323 K and 333 K). For

a given temperature, the adsorption equilibrium was obtained as a series of data pairs that

shows the maximum amount CO2 adsorbed (or CO2 loading in the unit of mmol CO2/ gm

of adsorbent) at the corresponding pressure of CO2 gas. The measured equilibrium data

for all adsorbents are given in Appendix A. The relationships between the adsorbed

amount of CO2 and equilibrium pressure, commonly known as isotherm curves, are given

in Figures 4.1 through 4.6 (for zeolite 13X, zeolite 5A, zeolite 4A, MSC-3R, GCA-830

and GCA-1240, respectively).

46
7.0
Amount of adsorbed CO2

6.0
5.0
(mmol /g)

4.0
3.0 293 K
303 K
2.0 313 K
323 K
1.0
333 K
0.0
0 10 20 30 40
Pressure (atm)

Figure 4.1: Isotherm curves of CO2 adsorption on zeolite 13X at different temperatures.

47
5.0
Amount of adsorbed CO2

4.0
(mmol/g)

3.0

2.0
293 K
303 K
1.0 313 K
333 K
0.0
0 10 20 30 40
Pressure (atm)

Figure 4.2: Isotherm curves of CO2 adsorption on zeolite 5A at different temperatures.

48
4.0
Amount of adsorbed CO2

3.0
(mmol /g)

2.0
293 K
303 K
1.0 323 K
333 K
0.0
0 10 20 30 40
Pressure (atm)

Figure 4.3: Isotherm curves of CO2 adsorption on zeolite 4A at different temperatures.

49
4.5
Amount of adsorbed CO2

3.0
(mmol/g)

293 K
1.5 303 K
313 K
323 K
333 K
0.0
0 10 20 30 40
Pressure (atm)

Figure 4.4: Isotherm curves of CO2 adsorption on MSC-3R at different temperatures.

50
10.5
Amount of adsorbed CO2

8.4
(mmol/g)

6.3

4.2 293 K
303 K
313 K
2.1 323 K
333 K
0.0
0 10 20 30 40
Pressure (atm)

Figure 4.5: Isotherm curves of CO2 adsorption on GCA-830 at different temperatures.

51
10.5
Amount of adsorbed CO2

8.4
(mmol/g)

6.3

4.2 293 K
303 K
313 K
2.1 323 K
333 K
0.0
0 10 20 30 40
Pressure (atm)

Figure 4.6: Isotherm curves of CO2 adsorption on GCA-1240 at different temperatures.

52
Note that the isotherm curves reported here were obtained from the experiments

using pure CO2 gas. These pure component isotherms can be used together with the real

adsorption solution theory (RAST) model for the design of a multi-component adsorption

system. From Figures 4.1 through 4.6, all isotherm curves exhibit common behaviour

regardless of temperature and adsorbent type (i.e., the amount of CO2 adsorbed on the

adsorbent increases very rapidly with the increase in equilibrium pressure over the low

pressure range, and it tends to stabilize as the pressure continues to increase). This

isotherm behaviour follows the type-I isotherm category according to IUPAC adsorption

isotherm classification (Keller et al., 2005), which indicates a monolayer adsorption

mechanism, commonly applied to microporous adsorbents (Thomas and Crittenden,

1998). The rapid increase in the isotherm curve demonstrates the proper range of

equilibrium pressure that promotes the majority of adsorption activity. This is important

information that can be used to identify both adsorption pressure and regeneration

pressure in the pressure swing adsorption process. The stabilized amount of CO2

adsorbed in the curve shows that all adsorption sites are occupied by CO2 gas, which

represents the capacity limit of the adsorbent at the particular temperature.

Figures 4.1 through 4.6 also reveal the typical behaviour showing the effect of

temperature on the CO2 adsorption capacity. That is, an increase in adsorption

temperature leads to a reduction in the amount of adsorbed CO2. Rising temperature

simply provides more internal energy to CO2 molecules in the gas phase. It should be

noted that the increasing energy allows gaseous molecules to diffuse at a greater rate, but,

at the same time, it reduces the chance for the CO2 to be restrained or trapped by fixed-

energy adsorption sites on the adsorbent surface.

53
Figure 4.7 shows a comparison of the CO2 adsorption isotherms among the

adsorbents tested in this study. The comparison was made at the lower-bound and upper-

bound temperatures: 293 K and 333 K, respectively. At 293 K, the activated carbons

(GCA-830 and GCA-1240) offered an adsorption capacity of 10.0 mmol/g, the highest

capacity among all the adsorbents tested. Zeolite 13X, zeolite 5A, and MSC-3R offered

CO2 adsorption capacities of 7.0, 4.7 and 4.2 mmol/g, respectively. Zeolite 4A provides

the lowest capacity of 3.8 mmol/g. The ranking of CO2 adsorption capacity, then, can be

written as GCA-1240 ˃ GCA-830 ˃ zeolite 13X ˃ zeolite 5A ˃ MSC-3R ˃ zeolite 4A.

This ranking remains unchanged as the temperature increases to 333 K. At the higher

temperature, the CO2 adsorption capacity for each adsorbent was reduced by 20 to 30 per

cent. The capacities at 333 K are 6.8 mmol/g for GCA-1240, 6.5 mmol/g for GCA-830,

5.2 mmol/g for zeolite 13X, 3.5 mmol/g for zeolite 5A, 3.1 mmol/g MSC-3R and 3.0

mmol/g for zeolite 4A.

In addition to the comparison of CO2 adsorption capacity, Figure 4.7 also reveals

the strength of interaction between CO2 molecules and adsorption sites for individual

adsorbents, which can be observed from the slope of the isotherm curves (during the

rapid increase). It appears that the zeolite-based adsorbents (zeolite 13X, zeolite 5A, and

zeolite 4A), exhibiting greater slope of the isotherm curves, have stronger adsorption sites

compare to the carbon molecular sieve (MSC-3R) and activated carbons (GCA-830 and

GCA-1240). The specific values of such interactions are reported in a later subsection

addressing the isosteric heat of adsorption for individual adsorbents.

54
10.5

Amount of adsorbed CO2


8.4

(mmol/g) 6.3

4.2

2.1 Zeolite 13x Zeolite 5A


Zeolite 4A MSC-3R
GCA-830 GCA-1240
0.0
0 10 20 30 40
Pressure (atm)
(a)

7.0
Amount of adsorbed CO2

5.6
(mmol/g)

4.2

2.8

Zeolite 13x Zeolite 5A


1.4
Zeolite 4A MSC-3R
GCA-830 GCA-1240
0.0
0 10 20 30 40
Pressure (atm)
(b)

Figure 4.7: Comparison of isotherm curves among the adsorbents tested (a) 293 K; (b)

333 K.

55
4.1.2 Isotherm correlations

To allow the use of the obtained adsorption isotherms for different purposes, the

equilibrium data were correlated into different isotherm models. These models are the

Langmuir, Toth, Sips, and Prausnitz equations presented previously in Chapter 2. Some

other isotherm models (Such as Volmer, Hill de Boer, Fowler-Guggenheim and Unilan)

which were discussed in Chapter 2 were not correlated with experimental isotherm data

in this thesis work because of the complexity of calculation method. Nonlinear regression

analysis was performed to determine the model parameters for individual adsorbents. The

obtained regression results for individual temperatures are given in Tables 4.1 through

4.5. The average percent deviation quantity, ∆q, or the adsorbed amounts were calculated

using the following formula (Morse et al., 2010):

∆q% = ∑ ⃒ ⃒ (4.1)

where h is the number of experimental data and Nexp and Ncal are the experimental and

calculated number of moles that are adsorbed by the adsorbent pellet. From Tables 4.1 to

4.5, it can be seen that the Sips model equation provides the best fit with the isotherm

curves for zeolite 13X, zeolite 5A, zeolite 4A, and carbon molecular sieve (MSC-3R),

which indicates that an adsorbed molecule can occupy more than one adsorption site

during the adsorption process (Morse et al., 2010). On the other hand, the Prausnitz

model equation provides the best fit for activated carbon (GCA-830 and GCA-1240),

which indicates that the adsorbent surface is not ideal but, rather, it is energetically

heterogeneous. It should be noted that both the Sips and Prausnitz models correlated the

obtained isotherm data well for temperatures ranging from 293 to 333 K. This enabled us

56
to develop a generic isotherm equation for each adsorbent that takes into account the

effect of adsorption temperature. The generic equations for corresponding adsorbent

materials are listed in Table 4.6.

57
Table 4.1: Regression parameters for different model equations at 293 K

Regression Parameters
2
Adsorbent Model a B t R ∆q%
Zeolite 13X Langmuir ‒ 3.40 ‒ 0.97 6.66
Toth ‒ 4.79 0.83 0.97 6.81
Sips ‒ 3.47 1.15 0.97 6.98
Prausnitz 28.29 6.44 0.02 0.98 5.23
Zeolite 5A Langmuir ‒ 4.38 ‒ 0.99 3.91
Toth ‒ 6.00 0.84 0.99 3.73
Sips ‒ 4.54 1.13 0.99 3.83
Prausnitz 23.81 4.39 0.01 0.99 3.07
Zeolite 4A Langmuir ‒ 1.76 ‒ 0.98 5.29
Toth ‒ 2.92 0.76 0.99 4.44
Sips ‒ 1.77 1.25 0.98 4.72
Prausnitz 9.18 3.19 0.05 0.99 3.42
MSC-3R Langmuir ‒ 0.83 ‒ 0.96 7.19
Toth ‒ 0.65 1.21 0.98 4.82
Sips ‒ 0.83 0.85 0.99 3.96
Prausnitz 3.11 4.67 -0.02 0.98 4.69
GCA-830 Langmuir ‒ 0.44 ‒ 0.96 7.40
Toth ‒ 0.31 1.28 0.99 3.93
Sips ‒ 0.42 0.85 0.99 3.48
Prausnitz 3.70 10.81 0.00 1.00 2.13
GCA-1240 Langmuir ‒ 0.43 ‒ 0.96 11.20
Toth ‒ 0.30 1.32 0.97 5.91
Sips ‒ 0.42 0.83 0.98 4.91
Prausnitz 3.59 12.01 -0.02 0.98 5.94

58
Table 4.2: Regression parameters for different model equations at 303 K

Regression Parameters
2 ∆q%
Adsorbent Model a B t R
Zeolite 13X Langmuir ‒ 1.90 ‒ 0.95 11.79
Toth ‒ 1.18 1.58 0.97 9.13
Sips ‒ 1.86 0.66 0.97 7.68
Prausnitz 9.52 7.07 -0.04 0.96 10.77
Zeolite 5A Langmuir ‒ 1.52 ‒ 0.93 8.11
Toth ‒ 0.82 1.68 0.97 5.63
Sips ‒ 1.32 0.65 0.98 4.56
Prausnitz 4.85 5.32 -0.06 0.96 7.00
Zeolite 4A Langmuir ‒ 0.85 ‒ 0.97 5.62
Toth ‒ 0.61 1.23 0.99 4.25
Sips ‒ 0.77 0.84 0.99 3.75
Prausnitz 2.53 3.96 -0.03 0.98 4.81
MSC-3R Langmuir ‒ 0.75 ‒ 0.95 9.18
Toth ‒ 0.40 1.63 1.00 2.49
Sips ‒ 0.65 0.71 1.00 1.93
Prausnitz 1.80 5.30 -0.09 1.00 2.74
GCA-830 Langmuir ‒ 0.36 ‒ 0.96 9.18
Toth ‒ 0.20 1.55 1.00 2.20
Sips ‒ 0.32 0.75 0.99 3.34
Prausnitz 2.04 11.92 -0.07 1.00 1.05
GCA-1240 Langmuir ‒ 0.36 ‒ 0.97 11.08
Toth ‒ 0.21 1.57 1.00 2.81
Sips ‒ 0.33 0.75 0.99 4.48
Prausnitz 2.17 12.71 -0.07 1.00 1.76

59
Table 4.3: Regression parameters for different model equations at 313 K

Regression Parameters
2 ∆q%
Adsorbent Model a B t R
Zeolite 13X Langmuir ‒ 1.65 ‒ 0.96 8.58
Toth ‒ 1.09 1.40 0.97 6.14
Sips ‒ 1.56 0.73 0.98 5.61
Prausnitz 7.58 6.30 -0.04 0.97 7.18
Zeolite 5A Langmuir ‒ 1.33 0.90 16.35
Toth ‒ 0.65 2.22 0.96 9.91
Sips ‒ 1.21 0.52 0.98 6.88
Prausnitz 3.64 5.71 -0.10 0.94 12.59
Zeolite 4A Langmuir ‒ ‒ ‒ ‒ ‒
Toth ‒ ‒ ‒ ‒ ‒
Sips ‒ ‒ ‒ ‒ ‒
Prausnitz ‒ ‒ ‒ ‒ ‒
MSC-3R Langmuir ‒ 0.68 0.95 7.70
Toth ‒ 0.36 1.64 1.00 1.27
Sips ‒ 0.59 0.70 1.00 2.06
Prausnitz 1.51 4.92 -0.09 1.00 0.94
GCA-830 Langmuir ‒ 0.31 ‒ 0.95 11.41
Toth ‒ 0.16 1.66 1.00 2.74
Sips ‒ 0.27 0.72 0.99 4.55
Prausnitz 1.58 11.73 -0.07 1.00 1.57
GCA-1240 Langmuir ‒ 0.33 ‒ 0.96 11.43
Toth ‒ 0.19 1.58 0.99 3.88
Sips ‒ 0.30 0.75 0.99 5.48
Prausnitz 1.84 10.94 -0.05 1.00 1.40

60
Table 4.4: Regression parameters for different model equations at 323 K

Regression Parameters
2 ∆q%
Adsorbent Model a B t R
Zeolite 13X Langmuir ‒ 1.44 ‒ 0.95 11.92
Toth ‒ 0.91 1.53 0.97 8.14
Sips ‒ 1.40 0.68 0.98 6.73
Prausnitz 6.14 6.27 -0.05 0.96 9.74
Zeolite 5A Langmuir ‒ ‒ ‒ ‒ ‒
Toth ‒ ‒ ‒ ‒ ‒
Sips ‒ ‒ ‒ ‒ ‒
Prausnitz ‒ ‒ ‒ ‒ ‒
Zeolite 4A Langmuir ‒ 0.53 ‒ 0.96 22.67
Toth ‒ 0.29 2.11 1.00 4.66
Sips ‒ 0.53 0.65 1.00 6.66
Prausnitz 1.14 6.43 -0.18 0.99 10.11
MSC-3R Langmuir ‒ 0.59 ‒ 0.94 11.97
Toth ‒ 0.28 1.84 1.00 2.38
Sips ‒ 0.49 0.66 1.00 2.05
Prausnitz 1.12 5.63 -0.13 1.00 2.94
GCA-830 Langmuir ‒ 0.30 ‒ 0.94 9.91
Toth ‒ 0.15 1.72 0.99 3.66
Sips ‒ 0.26 0.71 0.99 5.21
Prausnitz 1.37 10.75 -0.07 1.00 0.56
GCA-1240 Langmuir ‒ 0.29 ‒ 0.96 13.12
Toth ‒ 0.16 1.66 1.00 3.83
Sips ‒ 0.27 0.74 0.99 6.07
Prausnitz 1.37 10.67 -0.08 1.00 1.46

61
Table 4.5: Regression parameters for different model equations at 333 K

Regression Parameters
2 ∆q%
Adsorbent Model a B t R
Zeolite 13X Langmuir ‒ 1.35 ‒ 0.96 12.08
Toth ‒ 0.84 1.53 0.98 7.15
Sips ‒ 1.29 0.71 0.99 6.28
Prausnitz 5.41 6.36 -0.06 0.97 8.69
Zeolite 5A Langmuir ‒ 1.66 ‒ 0.95 6.33
Toth ‒ 0.92 1.61 0.98 3.41
Sips ‒ 1.42 0.67 0.99 2.61
Prausnitz 4.44 4.24 -0.05 0.97 4.53
Zeolite 4A Langmuir ‒ 0.34 ‒ 0.95 22.74
Toth ‒ 0.19 2.73 1.00 2.47
Sips ‒ 0.38 0.62 1.00 9.79
Prausnitz 0.67 9.41 -0.30 0.99 7.93
MSC-3R Langmuir ‒ 0.56 ‒ 0.93 11.21
Toth ‒ 0.26 1.90 1.00 2.42
Sips ‒ 0.47 0.64 1.00 1.90
Prausnitz 0.94 5.51 -0.15 1.00 2.64
GCA-830 Langmuir ‒ 0.27 ‒ 0.93 11.14
Toth ‒ 0.13 1.84 0.99 3.74
Sips ‒ 0.23 0.69 0.98 4.92
Prausnitz 1.05 12.49 -0.12 1.00 0.76
GCA-1240 Langmuir ‒ 0.27 ‒ 0.95 13.16
Toth ‒ 0.65 0.76 0.86 27.12
Sips ‒ 0.24 0.72 0.99 5.89
Prausnitz 1.11 12.00 -0.11 1.00 1.15

62
Table 4.6: Isotherm correlation equations for adsorbents tested

Adsorbent Model Temperature Generic Isotherm Equation


Equation Range (K)

q ∗T P 1.44 (71.43e −0.012 T )1.44


Zeolite 13x Sips 293 - 333 q=
1+ P 1.44 (71.43e −0.012 T )1.44

q ∗T P 1.64 (6.3783e −0.005T )1.64


Zeolite 5A Sips 293- 333 q=
1+ P 1.64 (6.378e −0.005 T )1.64

q ∗T P 1.43 (911.85e −0.023T )1.43


Zeolite 4A Sips 293 -333 q=
1+ P 1.43 (911.85e −0.023 T )1.43

q ∗T P 1.48 (8.473e −0.009 T )1.48


MSC-3R Sips 293 - 333 q=
1+ P 1.48 (8.473e −0.009 T )1.48

1 1 1
GCA-830 Prausnitz 293 - 333 q
=
1351.1e −0.021 T
+
P (58.13e −0.005 T )P −0.08

1 1 1
= +
GCA-1240 Prausnitz 293 - 333 q 2336.8e −0.023T P (175.85e −0.009 T )P −0.08

63
4.1.3 Isosteric heat of adsorption

According to the Clausius-Clapeyron equation (2.19), it is possible to analyze

isosteric heat of adsorption from a series of isotherm curves. The Clausius-Clapeyron

equation can be rewritten as:


lnP = - + constant (4.2)

Here, T is the adsorption temperature, P is the pressure, R is the gas constant, and ∆H is

the isosteric heat of adsorption. Analyzing Equation (4.2) helped identify the regeneration

energy as well as the interaction between adsorbent materials and gaseous molecules,

which provided important information on the level of heterogeneity of the solid surface.

As per the ideal Langmuir model (Langmuir, 1918), isosteric heat of adsorption should

be constant with the adsorbate or surface loading, suggesting that the solid surface is

energetically homogeneous. However, in the real system, the adsorbent’s surface is not

energetically homogeneous and the isosteric heat varies with the amount of adsorbate

retained on the surface (surface loading). Consequently, it is important to extract the

isosteric heat (∆H) from the slope of lnP ‒ (1/T) plot at a given adsorbent loading. Figure

4.8 show a series of these plots for individual adsorbents. The isosteric heats of

adsorption obtained from Figure 4.8 are listed in Table 4.7. The corresponding heat of

adsorption data are also presented in Figure 4.9 as a function of surface loading (or the

amount of CO2 adsorbed).

64
3.0 3.0

1.5 1.5

0.0 0.0

lnP
ln P
-1.5 -1.5

-3.0 -3.0

0.5 mmol/g 1 mmol/g 0.5 mmol/g 1 mmol/g


-4.5 -4.5
1.5 mmol/g 2 mmol/g 1.5 mmol/g 2 mmol/g
2.5 mmol/g 3 mmol/g 2.5 mmol/g 3 mmol/g
-6.0 -6.0
2.8 3.0 3.2 3.4 3.6 2.8 3.0 3.2 3.4 3.6
1/T x 10 3 (K-1) 1/T x 103 (K-1)
(a) (b)

3.0 3.0

1.5 1.5

0.0 0.0
lnP

-1.5 ln P -1.5

-3.0 -3.0
0.5 mmol/g 0.75 mmol/g 0.5 mmol/g 1 mmol/g
-4.5 1 mmol/g 1.25 mmol/g -4.5 1.5 mmol/g 2 mmol/g
1.5 mmol/g 1.75 mmol/g 2.5 mmol/g
-6.0 -6.0
2.8 3.0 3.2 3.4 3.6 2.8 3.0 3.2 3.4 3.6
1/T x 103 (K-1) 1/T x 103 (K-1)

(c) (d)

3.0 3.0

1.5 1.5

0.0 0.0
ln P
ln P

-1.5 -1.5

-3.0 -3.0
0.5 mmol/g 1 mmol/g 0.5 mmol/g 1 mmol/g
-4.5 1.5 mmol/g 2 mmol/g -4.5 1.5 mmol/g 2 mmol/g
2.5 mmol/g 3 mmol/g 2.5 mmol/g 3 mmol/g
-6.0 -6.0
2.8 3.0 3.2 3.4 3.6 2.8 3.0 3.2 3.4 3.6
1/T x 103 (K-1) 1/T x 103 (K-1)

(e) (f)

Figure 4.8: Plots of lnP versus 1/T for (a) zeolite 13x, (b) zeolite 5A, (c) zeolite 4A, (d)
molecular sieve carbon MSC-3R, (e) activated carbon GCA-830, and (f)
activated carbon GCA-1240

65
Table 4.7 Isosteric heat of CO2 adsorption on the adsorbents

Adsorbent Adsorbate loading ∆H Average ∆H


(mmol/g) (kJ/mol) (kJ/mol)
Zeolite 13X 0.50 40.10
1.00 39.70
1.50 39.30
39.23
2.00 39.00
2.50 38.70
3.00 38.60
Zeolite 5A 0.50 24.20
1.00 23.80
1.50 23.50
23.45
2.00 23.20
2.50 23.00
3.00 23.00
Zeolite 4A 0.50 46.40
0.75 45.00
1.00 43.60
42.78
1.25 42.20
1.50 40.60
1.75 38.90
MSC-3R 0.50 23.10
1.00 23.10
1.50 23.30 24.00
2.00 24.10
2.50 26.40
GCA-830 0.50 23.80
1.00 23.80
1.50 23.90
24.03
2.00 24.00
2.50 24.20
3.00 24.50
GCA-1240 0.50 23.40
1.00 22.40
1.50 22.10
22.73
2.00 22.00
2.50 22.20
3.00 24.30

66
It appears from Figure 4.9 that isosteric heat of CO2 adsorption on zeolite 13X,

zeolite 5A, and zeolite 4A decreases with the increasing adsorbate loading, indicating that

the interaction between adsorbent material and CO2 gas is rather strong at the initial stage

of adsorption and becomes weaker when the adsorption process continues. This

decreasing trend demonstrates that the regeneration of adsorbent material will be energy

intensive. Figure 4.9 also shows that the isosteric heat for activated carbons and the

carbon molecular sieve increases with the adsorbate loading. The reason for this

increasing trend is the Lennard-Jones potential energy of interaction between the

adsorbate molecules and adsorbent surface atoms, which depends on the position of the

adsorbate molecule and the size and shape of the pore of the adsorbent (Nguyen & Do,

1999). The increase in heat of adsorption with loading suggests that the adsorption is

reversible and the regeneration is not energy intensive. In addition, Figure 4.9 also

reveals the degree of difficulty regarding the adsorbent regeneration. The greater the heat

of adsorption, the more difficult and energy intensive the regeneration process. It is clear

that zeolite-based adsorbents would require more energy to regenerate compared to the

carbon based materials. Zeolite 4A clearly requires the highest regeneration energy

among all adsorbents tested.

67
45
Isosteric heat of adsorption

Zeolite 13X
40 Zeolite 5A
Zeolite 4A
MSC-3R
35
(kJ/mol)

GCA-830
GCA-1240
30

25

20
0.0 2.0 4.0 6.0
Amount of adsorbed CO2
(mmol/g)
Figure 4.9 Correlation plot of isosteric heat of adsorption versus adsorbate loading

68
4.2 CO2 adsorption kinetics

It is very important to measure the kinetics of an adsorbent so as to evaluate its

suitability for the adsorption application required. An adsorbent having high capacity

with slow kinetics or an adsorbent having fast kinetics with low capacity might not be the

ideal choice for industrial applications. Slow kinetics result in long residence time in the

adsorption column and low capacity requires more adsorbent to attain the desired product

(Do, 1998). The CO2 adsorption kinetics for the six adsorbents were measured through a

series of uptake rate experiments at temperatures from 293 to 333 K and pressures up to

11 atm.

Figures 4.10 through 4.15 show plots of the CO2 uptake rates of the six adsorbents

at different pressures. From the figures, it can be seen that, in all cases, CO2 was adsorbed

quickly on the adsorbent surface at the initial stage of adsorption, then the rate became

slower with time, and, finally, it reached adsorption equilibrium. It was also observed that

the CO2 adsorption processes at the higher pressures reached equilibrium to a greater

degree and, yet, in a shorter period compared to those at the lower pressures. This

suggests that an increase in pressure promotes CO2 adsorption rate.

69
7.0 7.0

Amount of adsorbed CO2


Amount of adsorbed CO2
6.0 6.0

5.0 5.0

(mmol / g)
(mmol / g) 4.0 4.0

3.0 3.0

2.0 2.0
293 K 293 K
1.0 313 K 1.0 313 K
333 K 333 K
0.0 0.0
0 150 300 450 600 750 0 150 300 450 600 750
Time (second) Time (second)
(a) (b)

7.0 7.0
Amount of adsorbed CO2

Amount of adsorbed CO2


6.0 6.0

5.0 5.0
(mmol / g)

(mmol / g)
4.0 4.0

3.0 3.0

2.0 2.0
293 K 293 K
1.0 313 K 1.0 313 K
333 K 333 K
0.0 0.0
0 150 300 450 600 750 0 150 300 450 600 750
Time (second) Time (second)

(c) (d)

7.0
Amount of adsorbed CO2

6.0

5.0
(mmol / g)

4.0

3.0

2.0
293 K
1.0 313 K
333 K
0.0
0 150 300 450 600 750
Time (second)
(e)

Figure 4.10: Plots of CO2 uptake for zeolite 13X at (a) 3.4 atm; (b) 5.4 atm; (c) 6.8 atm;
(d) 8.8 atm; (e) 10.9 atm.

70
5.0 5.0

Amount of adsorbed CO2


Amount of adsorbed CO2
3.8 3.8

(mmol / g)
(mmol / g)
2.5 2.5

1.3 1.3 293 K


293 K
313 K 313 K
333 K 333 K
0.0 0.0
0 150 300 450 600 750 0 150 300 450 600 750
Time (second) Time (second)
(a) (b)

5.0 5.0
Amount of adsorbed CO 2

Amount of adsorbed CO 2
3.8 3.8
(mmol / g)

(mmol / g)
2.5 2.5

1.3 293 K 1.3


293 K
313 K 313 K
333 K 333 K
0.0 0.0
0 150 300 450 600 750 0 150 300 450 600 750
Time (second) Time (second)
(c) (d)

5.0
Amount of adsorbed CO2

3.8
(mmol / g)

2.5

1.3 293 K
313 K
333 K
0.0
0 150 300 450 600 750
Time (second)
(e)

Figure 4.11: Plots of CO2 uptake for zeolite 5A at (a) 3.4 atm; (b) 5.4 atm; (c) 6.8 atm;
(d) 8.8 atm; (e) 10.9 atm.

71
4.0 4.0
Amount of adsorbed CO2

Amount of adsorbed CO2


3.0 3.0
(mmol / g)

(mmol / g)
2.0 2.0

1.0 293 K 1.0 293 K


313 K 313 K
333 K 333 K
0.0 0.0
0 150 300 450 600 750 900 0 150 300 450 600 750 900
Time (second) Time (second)
(a) (b)

4.0 4.0
Amount of adsorbed CO2

Amount of adsorbed CO2

3.0 3.0
(mmol / g)

(mmol / g)

2.0 2.0

1.0 293 K 1.0 293 K


313 K 313 K
333 K 333 K
0.0 0.0
0 150 300 450 600 750 900 0 150 300 450 600 750 900
Time (second) Time (second)
(c) (d)

Figure 4.12: Plots of CO2 uptake for zeolite 4A at (a) 3.4 atm; (b) 5.4 atm; (c) 6.8 atm;
(d) 8.8 atm.

72
4.5 4.5
Amount of adsorbed CO2

Amount of adsorbed CO2


(mmol / g) 3.6 3.6

(mmol / g)
2.7 2.7

1.8 1.8

293 K 293 K
0.9 0.9
313 K 313 K
333 K 333 K
0.0 0.0
0 170 340 510 680 850 0 170 340 510 680 850
Time (second) Time (second)
(a) (b)

4.5 4.5
Amount of adsorbed CO2

Amount of adsorbed CO2


3.6 3.6
(mmol / g)

2.7 (mmol / g) 2.7

1.8 1.8

293 K 293 K
0.9 0.9
313 K 313 K
333 K 333 K
0.0 0.0
0 170 340 510 680 850 0 170 340 510 680 850
Time (second) Time (second)
(c) (d)

4.5
Amount of adsorbed CO2

3.6
(mmol / g)

2.7

1.8

0.9 293 K
313 K
333 K
0.0
0 170 340 510 680 850
Time (second)
(e)

Figure 4.13: Plots of CO2 uptake for MSC-3R at (a) 3.4 atm; (b) 5.4 atm; (c) 6.8 atm;
(d) 8.8 atm; (e) 10.9 atm.

73
8.5 8.5
Amount of adsorbed CO2

Amount of adsorbed CO2


(mmol / g) 6.8 6.8

(mmol / g)
5.1 5.1

3.4 3.4

293 K 293 K
1.7 1.7
313 K 313 K
333 K 333 K
0.0 0.0
0 150 300 450 600 750 0 150 300 450 600 750
Time (second) Time (second)

(a) (b)

8.5 8.5
Amount of adsorbed CO 2

Amount of adsorbed CO 2
6.8 6.8
(mmol / g)

5.1 (mmol / g) 5.1

3.4 3.4

1.7 293 K 1.7 293 K


313 K 313 K
333 K 333 K
0.0 0.0
0 150 300 450 600 750 0 150 300 450 600 750
Time (second) Time (second)

(c) (d)

8.5
Amount of adsorbed CO2

6.8
(mmol / g)

5.1

3.4

1.7 293 K
313 K
333 K
0.0
0 150 300 450 600 750
Time (second)

(e)

Figure 4.14: Plots of CO2 uptake for GCA-830 at (a) 3.4 atm; (b) 5.4 atm; (c) 6.8 atm;
(d) 8.8 atm; (e) 10.9 atm.

74
9.0 9.0
Amount of adsorbed CO2

Amount of adsorbed CO2


7.5 7.5

6.0 6.0
(mmol / g)

(mmol / g)
4.5 4.5

3.0 3.0
293 K 293 K
1.5 313 K 1.5
313 K
333 K 333 K
0.0 0.0
0 150 300 450 600 750 0 150 300 450 600 750
Time (second) Time (second)
(a) (b)

9.0 9.0
Amount of adsorbed CO2

Amount of adsorbed CO2


7.5 7.5

6.0 6.0
(mmol / g)

(mmol / g)
4.5 4.5

3.0 3.0
293 K 293 K
1.5 313 K 1.5 313 K
333 K 333 K
0.0 0.0
0 150 300 450 600 750 0 150 300 450 600 750
Time (second) Time (second)

(c) (d)

9.0
Amount of adsorbed CO2

7.5

6.0
(mmol / g)

4.5

3.0
293 K
1.5 313 K
333 K
0.0
0 150 300 450 600 750
Time (second)
(e)

Figure 4.15: Plots of CO2 uptake for GCA-1240 at (a) 3.4 atm; (b) 5.4 atm; (c) 6.8 atm;
(d) 8.8 atm; (e) 10.9 atm.

75
4.2.1 Mass transfer coefficient for CO2 adsorption

It is essential to analyze the CO2 adsorption rate in terms of the mass transfer

coefficient for individual adsorbents for the purposes of comparison. In this study, the

mass transfer coefficient was analyzed from the plots of CO2 uptake rate presented

previously. The linear driving force (LDF) model representing the first order kinetics of

the adsorption process was used for the analysis. The LDF model was described in

Chapter 2. The integral form of the model was given earlier as Equation (2.18):


ln = - kt (2.18)

where k is the overall mass transfer coefficient presented as a function of adsorption

pressure and temperature. Here, the plots of CO2 uptake rate were translated into plots of

ln (1-q/q*) versus adsorption time (t) for the 70-80% of CO2 adsorption capacity of the

six commercial adsorbents, of which the slope represents the coefficient k at the

corresponding pressure and temperature. Figures 4.16 to 4.18 show plots of ln(1-q/q*)

versus time for all adsorbents tested at 293, 313, and 333 K, respectively. These plots

show linear behaviour, which confirms that the obtained experimental data follow the

LDF model and first order adsorption kinetics, validating the assumption of the data

analysis through LDF model. Mass transfer coefficients obtained from the LDF analysis

are reported in Table 4.8. From the table, it can be seen that, at a given pressure, mass

transfer coefficient increases with adsorption temperature. An increase in temperature

results in higher diffusion of CO2 molecules, providing greater mass transfer activities. It

can be seen from Figures 4.16 through 4.18 that zeolite 4A provided different mass

transfer behaviour, especially at the initial stage of adsorption, compared to the other

76
adsorbents tested. It seems that the mass transfer process took place relatively faster once

the adsorption started, and it was retarded as time progressed. This behaviour is probably

due to the micropore properties of zeolite 4A. The micropore diameter of zeolite 4A is

approximately 3.9 Å, and the molecular diameter of CO2 gas is 3.996 Å. Because of the

equivalent dimensions between adsorption pore and adsorbate molecule, the initial

adsorption occurred at a faster rate on the outer surface of zeolite 4A, and then the

diffusion through the pores took place later, at a slower rate. Table 4.8 show that the mass

transfer coefficient increased with an increased in pressure for the six commercial

adsorbents.

77
0.0 0.0
3.4 atm 3.4 atm
5.4 atm 5.4 atm
-0.2 -0.2
6.8 atm 6.8 atm

ln (1-q/q*)
ln (1-q/q*)
8.8 atm 8.8 atm
-0.4 10.9 atm -0.4 10.9 atm

-0.6 -0.6

-0.8 -0.8
0 150 300 450 0 150 300 450
Time (second) Time (second)

(a) Zeolite 13X (b) Zeolite 5A

0.0 0.0
3.4 atm 3.4 atm
5.4 atm
6.8 atm 5.4 atm
-0.2 -0.2
8.8 atm 6.8 atm
ln (1-q/q*)

ln (1-q/q*)
8.8 atm
-0.4 -0.4 10.9 atm

-0.6 -0.6

-0.8 -0.8
0 150 300 450 0 150 300 450
Time (second) Time (second)

(c) Zeolite 4A (d) MSC-3R

0.0 0.0
3.4 atm 3.4 atm
5.4 atm
5.4 atm
-0.2 6.8 atm -0.2
6.8 atm
8.8 atm
ln (1-q/q*)
ln (1-q/q*)

10.9 atm 8.8 atm


-0.4 -0.4 10.9 atm

-0.6 -0.6

-0.8 -0.8
0 150 300 450 0 150 300 450
Time (second) Time (second)

(e) GCA-830 (f) GCA-1240

Figure 4.16: Linear plot of ln (1- ∗) versus time for CO2 adsorption at 293 K and
different pressures

78
0.0 0.0
3.4 atm 3.4 atm
5.4 atm 5.4 atm
-0.2 -0.2
6.8 atm 6.8 atm

ln (1-q/q*)
ln (1-q/q*)
8.8 atm 8.8 atm
-0.4 10.9 atm -0.4 10.9 atm

-0.6 -0.6

-0.8 -0.8
0 150 300 450 0 150 300 450
Time (second) Time (second)

(a) Zeolite 13X (b) Zeolite 5A

0.0 0.0
3.4 atm 3.4 atm
5.4 atm
6.8 atm 5.4 atm
-0.2 -0.2
8.8 atm 6.8 atm
ln (1-q/q*)

ln (1-q/q*)
8.8 atm
-0.4 -0.4 10.9 atm

-0.6 -0.6

-0.8 -0.8
0 150 300 450 0 150 300 450
Time (second) Time (second)

(c) Zeolite 4A (d) MSC-3R

0.0 0.0
3.4 atm 3.4 atm
5.4 atm
5.4 atm
-0.2 6.8 atm -0.2
6.8 atm
8.8 atm
ln (1-q/q*)
ln (1-q/q*)

10.9 atm 8.8 atm


-0.4 -0.4 10.9 atm

-0.6 -0.6

-0.8 -0.8
0 150 300 450 0 150 300 450
Time (second) Time (second)
(f) GCA-1240
(e) GCA-830

Figure 4.17: Linear plot of ln (1- ∗) versus time for CO2 adsorption at 313 K and
different pressures

79
0.0 0.0
3.4 atm 3.4 atm
5.4 atm 5.4 atm
-0.2 -0.2
6.8 atm 6.8 atm

ln (1-q/q*)
ln (1-q/q*)
8.8 atm 8.8 atm
-0.4 10.9 atm -0.4 10.9 atm

-0.6 -0.6

-0.8 -0.8
0 150 300 450 0 150 300 450
Time (second) Time (second)

(a) Zeolite 13X (b) Zeolite 5A


0.0 0.0
3.4 atm 3.4 atm
5.4 atm 5.4 atm
-0.2 6.8 atm -0.2
6.8 atm
ln (1-q/q*)

8.8 atm

ln (1-q/q*)
8.8 atm
-0.4 -0.4 10.9 atm

-0.6 -0.6

-0.8 -0.8
0 150 300 450 0 150 300 450
Time (second) Time (second)

(c) Zeolite 4A (d) MSC-3R

0.0 0.0
3.4 atm 3.4 atm
5.4 atm
5.4 atm
-0.2 6.8 atm -0.2
6.8 atm
8.8 atm
ln (1-q/q*)
ln (1-q/q*)

10.9 atm 8.8 atm


-0.4 -0.4 10.9 atm

-0.6 -0.6

-0.8 -0.8
0 150 300 450 0 150 300 450
Time (second) Time (second)

(e) GCA-830 (f) GCA-1240

Figure 4.18: Linear plot of ln (1- ∗) versus time for CO2 adsorption at 333 K and
different pressures

80
Table 4.8: Mass transfer coefficients of CO2 adsorption on the tested adsorbents
3
Pressure Temperature Mass transfer coefficients, k x 10 (1/sec)
(atm) (K) Zeolite 13X Zeolite 5A Zeolite 4A MSC-3R GCA-830 GCA-1240
3.41 293 1.79 1.73 0.77 1.79 1.39 1.46
313 2.00 2.01 0.99 2.13 1.61 1.66
333 2.23 2.23 1.56 2.31 1.82 1.90
5.44 293 2.84 2.14 0.92 2.22 2.86 2.95
313 3.00 2.39 1.19 2.46 3.09 3.16
333 3.20 2.65 1.74 2.72 3.29 3.34
6.81 293 3.60 2.74 1.31 3.04 4.46 4.68
313 3.72 2.94 1.71 3.24 4.61 4.80
333 3.93 3.20 1.95 3.48 4.72 4.93
8.84 293 5.34 3.52 1.65 3.97 5.53 6.00
313 5.57 3.82 1.90 4.18 5.63 6.13
333 5.65 3.98 2.10 4.38 5.72 6.20
10.88 293 6.29 5.16 ‒ 5.13 6.86 7.15
313 6.41 5.35 ‒ 5.27 6.93 7.22
333 6.50 5.65 ‒ 5.44 7.07 7.31

81
4.2.2 Mass transfer coefficient and activation energy

From the previously shown Table 4.8, it can be inferred that an increase in

adsorption temperature results in an increase in mass transfer coefficient k regardless of

adsorbent material. This temperature effect can be demonstrated through the Arrhenius

equation shown earlier in Chapter 2:

lnk = - + lnA (2.21)

Here, the activation energy (Ea) can be extracted from the slope of the plot between lnk

and the reciprocal of temperature ( ). Figure 4.19 shows the linear plots of lnk versus ( )

based on the experimental data obtained in this study for zeolite 13x, zeolite 5A, zeolite

4A, MSC 3R, GCA-830, and GCA-1240. Table 4.9 summarizes the activation energy and

frequency factor A that were analyzed from Figure 4.19. It is clear that the activation

energy decreases with pressure, which suggests strong adsorbate-adsorbent interaction

potential at high pressure, and considering the heterogeneity of the adsorbent surface, the

CO2 adsorption at low pressure occurs mostly on adsorption sites with strong energy

barriers while adsorption at high pressure takes place on the sites with weak energy

barriers. Table 4.9 also shows that the frequency factor A increases with increasing

pressure, thereby promoting increased collisions between adsorbate molecules and

adsorbent materials, which results in a reduction in activation energy for the adsorption

process. Figure 4.20 shows the significance of the pressure effect on the activated energy

and the frequency factor. The effect of pressure was then correlated and included in the

Arrhenius-based mass transfer equation shown below from Equation (4.3) to Equation

(4.8):

82
. . .
Zeolite 13X: k = (0.01P2 + 0.48P + 0.67) exp (‒ ) (4.3)

. . .
Zeolite 5A: k = (0.05P2 ‒ 0.28P + 2.94) exp (‒ ) (4.4)

. – . .
Zeolite 4A: k = (‒ 0.01P2 + 0.12P + 1.73) exp (‒ ) (4.5)

. – . .
MSC-3R: k = (0.03P2 + 0.05P + 2.13) exp (− ) (4.6)

. – . .
GCA-830: k = (‒ 0.02P2 + 0.96P - 0.96) exp (− ) (4.7)

. – . .
GCA-1240: k = (‒ 0.03P2 + 1.10P - 1.36) exp (‒ ) (4.8)

Figure 4.21 shows a comparison of activation energy amongst the adsorbents tested

in this study. Regardless of pressure, the activation energy of zeolite 4A was much higher

than the activation energy of the other adsorbents tested. At pressures below 3.4 atm,

zeolite 5A and MSC-3R offered lower activation energies than zeolite 13X and the

activated carbons. However, at the higher pressure range, zeolite 13X and the activated

carbons required less activation energy than zeolite 5A and the carbon molecular sieve.

This comparison suggests that, for high pressure applications, zeolite 13X and the

activated carbons are the more suitable adsorbents for fast removal of CO2 from

industrial gas streams. However, for a specific application of CO2 removal from high-

pressure natural gas, zeolite 13X appears to be the best adsorbent because it tends to

adsorb less hydrocarbons compared to the activated carbons.

83
-4.0 -4.0

-5.0 -5.0

ln k
ln k

-6.0 -6.0

-7.0 -7.0
3.4 atm 5.4 atm 3.4 atm 5.4 atm
6.8 atm 8.8 atm 6.8 atm 8.8 atm
10.9 atm 10.9 atm
-8.0 -8.0
2.8 3.0 3.2 3.4 3.6 2.8 3.0 3.2 3.4 3.6
1/T x 10 3 (K-1) 1/T x 10 3 (K-1)

(a) (b)
-4.0 -4.0

-5.0 -5.0
ln k

ln k
-6.0 -6.0

-7.0 -7.0 3.4 atm 5.4 atm


3.4 atm 5.4 atm 6.8 atm 8.8 atm
6.8 atm 8.8 atm 10.9 atm
-8.0 -8.0
2.8 3.0 3.2 3.4 3.6 2.8 3.0 3.2 3.4 3.6
1/T x 10 3 (K-1) 1/T x 103 (K-1)

(c) (d)
-4.0 -4.0

-5.0 -5.0
ln k
ln k

-6.0 -6.0

-7.0 -7.0
3.4 atm 5.4 atm 3.4 atm 5.4 atm
6.8 atm 8.8 atm 6.8 atm 8.8 atm
10.9 atm 10.9 atm
-8.0 -8.0
2.8 3.0 3.2 3.4 3.6 2.8 3.0 3.2 3.4 3.6
1/T x 103 (K-1) 1/T x 10 3 (K-1)

(e) (f)

Figure 4.19: Linear plot of lnk versus ( ) for CO2 adsorption activation energy on (a)

zeolite 13X; (b) zeolite 5A; (c) zeolite 4A; (d) MSC-3R; (e) GCA-830; (f)

GCA-1240

84
Table 4.9: Activation energy (Ea) and frequency factor (A) for the tested adsorbents

3
Adsorbent Pressure Ea A x 10
(atm) (kJ/mol) (1/sec)
Zeolite 13X 0.50 10.26 ‒
3.40 4.46 2.48
5.44 2.41 3.38
6.80 1.77 4.08
8.84 1.16 5.84
10.88 0.67 6.61
Zeolite 5A 0.50 8.08 ‒
3.40 5.17 2.54
5.44 4.34 2.94
6.80 3.14 3.44
8.84 2.51 4.25
10.88 1.83 5.88
Zeolite 4A 0.50 29.94 ‒
3.40 14.25 2.11
5.44 12.82 2.31
6.80 8.12 2.42
8.84 4.91 2.58
MSC-3R 0.50 7.80 ‒
3.40 5.21 2.66
5.44 4.12 3.00
6.80 2.74 3.71
8.84 1.99 4.60
10.88 1.19 5.59
GCA-830 0.50 12.04 ‒
3.40 5.47 2.08
5.44 2.85 3.53
6.80 1.15 4.86
8.84 0.69 5.82
10.88 0.61 7.16
GCA-1240 0.50 11.17 ‒
3.40 5.34 2.15
5.44 2.52 3.55
6.80 1.05 5.05
8.84 0.67 6.31
10.88 0.45 7.38

85
35 12 35 12
Activation energy Activation energy

Frequency factor x10 3


30 30 Frequency factor

Frequency factor x 10 3
Frequency factor

Activation energy
Activation energy
9 9
25 25

( kJ/ mol )

( kJ/ mol )
(1/sec)

(1/sec)
20 20
6 6
15 15

10 10
3 3
5 5

0 0 0 0
0 3 6 9 12 15 0 3 6 9 12 15
Pressure (atm) Pressure (atm)
(a) (b)

35 12 35 12
Activation energy Activation energy
30 30

Frequency factor x 10 3

Frequency factor x 10 3
Frequency factor Frequency factor
Activation energy

Activation energy
9 9
25 25

( kJ/ mol )
( kJ/ mol)

(1/sec)

(1/sec)
20 20
6 6
15 15

10 10
3 3
5 5

0 0 0 0
0 3 6 9 12 15 0 3 6 9 12 15
Pressure (atm) Pressure (atm)
(c) (d)

35 12 35 12
Activation energy Activation energy
30 30
Frequency factor x 10 3

Frequency factor x 10 3
Frequency factor Frequency factor
Activation energy

Activation energy

9 9
25 25
( kJ/ mol )

( kJ/ mol)
(1/sec)

(1/sec)
20 20
6 6
15 15

10 10
3 3
5 5

0 0 0 0
0 3 6 9 12 15 0 3 6 9 12 15
Pressure (atm) Pressure (atm)

(e) (f)
Figure 4.20: Effect of pressure on activation energy and frequency factor for (a) zeolite

13X; (b) zeolite 5A; (c) zeolite 4A; (d) MSC-3R; (e) GCA-830; (f) GCA-

1240

86
15
Zeolite 13X
Zeolite 5A
12
Activation energy

Zeolite 4A
MSC-3R
( kJ/ mol )

9 GCA-830
GCA-1240
6

0
0 3 6 9 12 15
Pressure (atm)

Figure 4.21: Comparison of activation energy of six different adsorbents

87
5. CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE WORK

5.1 Conclusions
This thesis has significantly extended the knowledge of carbon dioxide (CO2)

separation from natural gas using a pressure swing adsorption process. A large set of

experimental data on CO2 adsorption equilibrium and kinetics was produced for six

commercial adsorbents including zeolite 13X, zeolite 5A, zeolite 4A, a carbon molecular

sieve (MSC-3R), and two activated carbons (GCA-830 & GCA-1240). The CO2

adsorption experiments were conducted using the volumetric method at different

temperatures and pressures. From this study, the following conclusions can be drawn:

 The CO2 adsorption isotherm obtained in this study followed general gas

adsorption behaviour, demonstrating that the CO2 adsorption capacity increases

with increasing pressure and decreases with increasing temperature. The

adsorption isotherm follows a type-I isotherm classification according to IUPAC,

representing a monolayer adsorption mechanism. Among the six commercial

adsorbents tested, activated carbon GCA-1240 offers the highest adsorption

capacity, and zeolite 4A provides the lowest capacity for temperatures ranging

from 293 to 333 K and pressures up to 35 atm.

 The equilibrium data of CO2 adsorption were correlated to fit with different model

equations (i.e., the Langmuir, Toth, Sips, and Prausnitz equations). It was found

that the Sips model showed the best fit with the equilibrium data for zeolite 13X,

zeolte 5A, zeolite 4A, and the carbon molecular sieve (MSC-3R) while the

Prausnitz model provided excellent fit with the data for the activated carbons

88
(GCA-830 and GCA-1240). The correlations for individual adsorbents were

reported as a function of temperature and pressure.

 Isosteric heat of adsorption of CO2 adsorption on the six commercial adsorbents

was calculated using the Clausius-Clapeyron equation. It was found that the

isosteric heat of adsorption varied with CO2 loading, indicating the heterogeneity

of the adsorbent surface. The heat of adsorption of zeolite 13X, zeolite 5A, and

zeolite 4A decreased with increasing CO2 loading while the heat of adsorption of

the activated carbons and carbon molecular sieve increased with increasing CO2

loading.

 A comprehensive set of mass transfer coefficients for CO2 adsorption on the six

adsorbents was produced for a temperature range of 293 to 333 K and pressure up

to 11 atm. Among the tested adsorbents, zeolite 4A showed the lowest mass

transfer coefficient while activated carbon GCA-1240 offered the highest

coefficient. Regardless of adsorbents, the effect of temperature on the mass

transfer coefficient followed the Arrhenius equation and the pressure exhibited a

non-linear effect on the mass transfer coefficient.

 Activation energy of CO2 adsorption on the six adsorbents was analyzed using the

Arrhenius equation. This information revealed the amount of energy needed for

regenerating of the CO2 saturated adsorbent. It was found that the energy

decreased with pressure regardless of adsorbent. Among the six adsorbents, the

carbon molecular sieve had the lowest activation energy below pressure 3.4 atm

whereas zeolite 4A required the highest activation energy.

89
5.2 Recommendations for future work

To improve selection of the best adsorbent for the removal of CO2 from natural gas,

the following future research activities are recommended:

 The kinetic and equilibrium selectivity for CO2 adsorption compared to other

gases (such as methane and ethane) should be investigated in the future for the six

adsorbents as selectivity is an important criterion to choose an adsorbent for the

separation of a gas from the mixture of gases.

 The performance of combined adsorption-desorption cycles should be examined

to identify the proper cycling operation of the adsorbents which will provide the

necessary information regarding how long the adsorbent will show proper

separation efficiency for the separation of desired gas component.

 Cost analysis (capital and operating expenditures) should be performed for each

adsorbent to be used in the natural gas purification applications so that it would be

possible to select an adsorbent from technological as well as economical point of

view.

90
REFERENCES

1. Ahn, H., Moon, J., Hyun, S. & Lee, C. (2004). Diffusion Mechanism of Carbon

Dioxide in Zeolite 4A and CaX Pellets. Adsorption, 10, 111–128.

2. Amoros, D. C., Alcaniz-Monge, J. & Linares-Solano, A. (1996). Characterization

of activated carbon fibers by CO2 adsorption. Langmuir, 12, 2820-2824.

3. Amoros, D. C., Alcaniz-Monge, J., Casa-Lillo, M. A. & Linares-Solano, A.

(1998). CO2 as an adsorptive to characterize carbon molecular sieves and

activated carbons. Langmuir, 14, 4589-4596.

4. Bae, Y. S. & Lee, C. H. (2005). Sorption kinetics of eight gases on a carbon

molecular sieve at elevated pressure. Carbon 43, 95–107.

5. Calleja, G., Jimenez, A., Pau, J., Dominguez, L. & Perez, P. (1994).

Multicomponent adsorption equilibrium of ethylene, propane, propylene and CO2

on 13X zeolite. Gas Separation & Purification, 8, 4.

6. Castello, D. L., Amoros, D. C. & Solano, A. L. (2004). Usefulness of CO2

adsorption at 273 K for the characterization of porous carbons. Carbon 42, 1231–

1236.

7. Castello, D. L., Monge, J. A., Amoros, D. C., Solano, A. L, Zhu, W., Kapteijn, F.

& Moulijn, J.A. (2005). Adsorption properties of carbon molecular sieves

prepared from an activated carbon by pitch pyrolysis. Carbon, 43, 1643–1651.

8. Cavenati, S., Grande, C.A. & Rodrigues, A.E. (2004). Adsorption Equilibrium of

Methane, Carbon Dioxide, and Nitrogen on Zeolite 13X at High Pressures. J.

Chem. Eng. Data, 49, 1095-1101.

9. Chen, Y. D., Ritter, J. A. & Yang, R. T. (1990). Non-ideal adsorption from

91
multicomponent gas mixtures at elevated pressures on a 5A molecular sieve.

Chemical Engineering Science, 45, 2877-2894.

10. Chue, K. T., Kim, J. N., Yoo, Y. J. & Cho, S. H. (1995). Comparison of Activated

Carbon and Zeolite 13X for CO2 Recovery from Flue Gas by Pressure Swing

Adsorption. Ind. Eng. Chem. Res., 34, 591-598.

11. Costa, E., Sotelo, J.L., Calleja, G. & Marron, C. (1981). Adsorption of Binary and

Ternary Hydrocarbon Gas Mixtures on Activated Carbon: Experimental

Determination and Theoretical Prediction of the Ternary Equilibrium Data.

AIChE J., 27, 5–12.

12. Costa, E., Calleja, G., Marron, C., Jimenez, A. & Pau, J. (1989). Equilibrium

Adsorption of Methane, Ethane, Ethylene, and Propylene and Their Mixtures on

Activated Carbon.J Chem Eng Data, 34, 156-160.

13. Costa, E., Calleja, G., Jlmenez, A. & Pau, J. (1991). Adsorption Equilibrium of

Ethylene, Propane, Propylene, CarbonDioxide, and Their Mixtures on 13X

Zeolite. J. Chem. Eng. Data, 36, 218-224.

14. De Boer, J. H. (1953). The Dynamical Character of Adsorption. Oxford, The

Clarendon Press.

15. Do, D.D. (1998). Adsorption Analysis: Equilibria and Kinetics. London, Imperial

College Press.

16. Drage, T. C., Blackman, J. M., Pevida, C. & Snape, C. E. (2009). Evaluation of

activated carbon adsorbents for CO2 capture in gasification. Energy & Fuels, 23,

2790–2796.

92
17. Dreisbach, F., Staudt, R. & Keller, J.U. (1999).High pressure adsorption data of

methane, nitrogen, carbon dioxide and their binary and ternary mixtures on

activated carbon. Adsorption, 5, 215–227.

18. Eagan, J. D. & Anderson, R. B. (1975). Kinetics and Equilibrium of Adsorption

on 4A Zeolite. Journal of Colloid and Interface Science, Vol. 50, No. 3, 419-433.

19. Farooq, S., Qinglin, H. & Karimi I. A. (2002). Identification of Transport

Mechanism in Adsorbent Micropores from Column Dynamics. Ind. Eng. Chem.

Res., 41, 1098-1106.

20. Fowler, R. H. & Guggenheim, E. A. (1965). Statistical Thermodynamics.

Cambridge, Cambridge University Press.

21. Freundlich, H. (1907). Ueber die adsorption in loesungen. Z. Phys. Chem. 57,

385–470.

22. Fridleifsson, I. B. (2003).Status of geothermal energy amongst the world’s energy

sources. Geothermics. 32, 379–388.

23. Glueckauf, E. & Coates, J. J. (1947).Theory of chromatography. Part IV. The

influence of incomplete equilibrium on the front boundary of chromatograms and

on the effectiveness of separation. J. Chem. Soc., 1315.

24. Guo, B., Chang, L. & Xie, K. (2006). Adsorption of carbon dioxide on activated

carbon. Journal of Natural Gas Chemistry, 15, 223-229.

25. Harlick, P.J.E. & Tezel, F. H. (2004). An experimental adsorbent screening study

for CO2 removal from N2. Microporous and Mesoporous Materials, 76, 71–79.

26. Haynes, J. R. & Sharma, H. W. (1975). The determination of effective diffusivity

by gas chromatography time domain solutions. Chem. Eng. Sci., 30, 955.

93
27. Hill, T.L. (1946). Statistical Mechanics of Multimolecular Adsorption II.

Localized and Mobile Adsorption and Absorption. Journal of Chemical Physics,

14, 441-453.

28. Hill, T. L. (1949). Statistical Mechanics of Adsorption. V. Thermodynamics and

Heat of Adsorption. J. Chem. Phys., 17, 520-535.

29. Jayaraman, A., Chiao, A. S., Padin, J., Yang, R. T. & Munson, C. L. (2002).

Kinetic separation of methane/carbon dioxide by molecular sieve carbons. Sep.

Sci. Technol., 37(11), 2505–2528.

30. Keller, J.U. (1996). In Fundamentals of Adsorption V, edited by M.D. LeVan.

Boston, Massachusetts, Kluwer Academic Publishers.

31. Keller, J. U. & Staudt, R. (2005). Gas Adsorption Equilibria: Experimental

Methods and Adsorptive Isotherms. Boston, Springer Science and Business

Media, Inc.

32. Key World Energy Statistics, International Energy Agency, 2011, web site:

http://www.iea.org/publications/freepublications/publication/key_world_energy_s

tats-1.pdf

33. Kidnay, A.J. & Parrish, W. R. (2006). Fundamentals of Natural Gas Processing.

New York , Taylor and Francis Group, LLC.

34. Kikkinides, E. S. & Yang, R. T. (1993). Concentration and Recovery of CO2 from

Flue Gas by Pressure Swing Adsorption. Ind. Eng. Chem. Res., 32, 2714-2720.

35. Kim, W. G., Yang, J., Han, S., Cho, C., Leer, C. H. & Lee, H. (1995).

Experimental and theoretical study on H2/CO2 separation by a five-step one-

column psa process. Korean J. of Chem. Eng., 12(5), 503-511.

94
36. Kohl, A.L. & Nielsen, R.B. (1997). Gas Purification, 5th edition. Houston, Texas:

Gulf Publishing Company.

37. Ko, D., Siriwardane, R., & Biegler, L. T. (2003). Optimization of a Pressure-

Swing Adsorption Process Using Zeolite 13X for CO2 Sequestration. Ind. Eng.

Chem. Res., 42, 339-348.

38. Langmuir, I. (1918). The adsorption of gases on plane surfaces of glass, mica and

platinum. J. Am. Chem. Soc., 40, 1361.

39. Lee, J. S., Kim, J. H., Kim, J. T., Suh, J. K., Lee, J. M. & Lee, C. H.(2002).

Adsorption equilibria of CO2 on zeolite 13X and zeolite X/Activated carbon

composite. J. Chem. Eng. Data, 47, 1237-1242.

40. Li, G., Xiao, P., Webley, P., Zhang, J., Singh, R. & Marshall, M. (2008). Capture

of CO2 from high humidity flue gas by vacuum swing adsorption with zeolite

13X. Adsorption, 14, 415–422.

41. Liu, Z., Grande, C. A., Li, P., Yu, J. & Rodrigues, A.E. (2011). Adsorption and

Desorption of Carbon Dioxide and Nitrogen on Zeolite 5A. Separation Science

and Technology, 46, 434–451.

42. Markham, E. C. & Benton, A. F. (1931). The adsorption of gas mixtures by silica.

J. Am. Chem. Soc, 53, 497-507.

43. Merel, J., Clausse, M. & Meunier, F. (2008). Experimental Investigation on CO2

Post-CombustionCapture by Indirect Thermal Swing Adsorption Using 13X and

5A Zeolites. Ind. Eng. Chem. Res., 47, 209-215.

95
44. Millward, A. R. & Yaghi, O. M. (2005). Metal-organic frameworks with

exceptionally high capacity for storage of carbon dioxide at room temperature. J.

AM. Chem. Soc., 127, 17998-17999.

45. Mochida, I., Yatsunami, S., Kawabuchi, Y. & Nakayama, Y. (1995). Influence of

heat-treatment on the selective adsorption of CO2 in a model natural gas over

molecular sieve carbons. Carbon, 33(11), 1611-1619.

46. Mokhatab, S., W.A. Poe & J.G. Speight, 2006. Handbook of Natural Gas

Transmission and Processing. USA: Elsevier Inc., ISBN-10: 0-7506-7776-7.

47. Morse, G., Jones, R., Thibault, J. & Tezel, F. H. (2010). Neural network

modeling of adsorption isotherms. Adsorption, 17(2), 303-309.

48. Myers, A. L. & Prausnitz, J. M. (1965) Thermodynamics of mixed-gas

adsorption. AIChE J., 11, 121.

49. Nguyen, C. & Do, D. D. (1999). Adsorption of supercritical gases in porous

media: determination of micropore size distribution. J. Phys. Chem. B, 103, 6900-

6908.

50. Pakseresht, S., Kazemeini, M. & Akbarnejad, M. M. (2002). Equilibrium

isotherms for CO, CO2, CH4 and C2H4 on the 5A molecular sieve by a simple

volumetric apparatus. Separation and Purification Technology, 28, 53–60.

51. Perez, A. R. & Armenta, G. A. (2010). Adsorption Kinetics and Equilibria of

Carbon Dioxide, Ethylene, and Ethane on 4A (CECA) Zeolite. J. Chem. Eng.

Data, 55, 3625–3630.

52. Quadrelli, R. & Peterson, S. (2007). The energy–climate challenge: Recent trends

in CO2 emissions from fuel combustion. Energy Policy. 35, 5938–5952.

96
53. Radke, C. J. & Prausnitz, J. M. (1972). Adsorption of Organic Solutes from

Dilute Aqueous Solution on Activated Carbon. Ind. Eng. Chem. Fundam., 11 (4),

445.

54. Reid, C. R. & Thomas, K. M. (1999). Adsorption of gases on a carbon molecular

sieve used for air separation: linear adsorptives as probes for kinetic selectivity.

Langmuir, 15, 3206-3218.

55. Rodil, S.V., Navarrete, R., Denoyel, R., Albiniak, A., Paredes, J. I., Alonso, A. M.

& Tascon, J.M.D. (2005). Carbon molecular sieve cloths prepared by chemical

vapour deposition of methane for separation of gas mixtures. Microporous and

Mesoporous Materials, 77, 109–118.

56. Rutherford, S.W. & Do, D.D. (2000). Adsorption dynamics of carbon dioxide on

a carbon molecular sieve 5A. Carbon, 38, 1339–1350.

57. Rutherford, S. W., Nguyen, C., Coons, J. E. & Do, D. D. (2003). Characterization

of Carbon Molecular Sieves Using Methane and Carbon Dioxide as Adsorptive

Probes. Langmuir 19, 8335-8342.

58. Ruthven, D.M. (1984). Principles of Adsorption and Adsorption Process.

NewYork, NY: John Wiley and Sons, Inc.

59. Saha, D. E., Bao, Z., Jia, F. & Deng, S. (2010). Adsorption of CO2, CH4, N2O,

and N2 on MOF-5, MOF-177, and Zeolite 5A. Environ. Sci. Technol., 44, 1820–

1826.

60. Salehi, E., Taghikhani, V., Ghotbi, C., Lay, E. N. & Shojaei, A.

(2007).Theoretical and Experimental Study on the Adsorption and Desorption of

97
Methane by Granular Activated Carbon at 25 ℃. Journal of Natural Gas

Chemistry, 16, 415–422.

61. Sips, R. (1948). On the structure of a catalyst surface. J. Chem. Phys., 16, 490-

495.

62. Siriwardane, R. V., Shen, M. S., Fisher, E. P. & Poston, J. A. (2001). Adsorption

of CO2 on Molecular Sieves and Activated Carbon. Energy & Fuels, 15, 279-284.

63. Siriwardane, R. V., Shen, M. S. & Fisher, E. P. (2005). Adsorption of CO2 on

Zeolites at Moderate Temperatures. Energy & Fuels, 19, 1153-1159.

64. Sudibandriyo, M., Pan, Z., Fitzgerald, J. E., Jr., R. L. R. & Gasem, K. A. M.

(2003). Adsorption of methane, nitrogen, carbon dioxide, and their binary

mixtures on dry activated carbon at 318.2 K and pressures up to 13.6 MPa.

Langmuir, 19, 5323-5331.

65. Tagliabue, M., Farrusseng, D., Valencia, S., Aguado, S., Ravon, U., Rizzo, C.,

Corma, A., & Mirodatos, C. (2009). Natural gas treating by selective adsorption:

Material science and chemical engineering interplay. Chemical Engineering

Journal, 155, 553–566.

66. Talu, O. & Zwiebel, I. (1986). "Multicomponent adsorption equilibria of nonideal

mixtures; AIChE J., 32, 1263-1276.

67. Tan, J. S. & Ani, F. N. (2004). Carbon molecular sieves produced from oil palm

shell for air separation. Separation and Purification Technology, 35, 47–54.

68. Thomas, W.J. & Crittenden, B. (1998). Adsorption Technology and Design.

London: Reed educational and professional publishing Ltd.

98
69. Tlili, N., Grevillot, G., & Vallieres, C. (2009). Carbon dioxide capture and

recovery by means of TSA and/or VSA. International Journal of Greenhouse Gas

Control 3, 519–527.

70. Toth, J. (1971). State equations of the solid gas interface layer. Acta Chim. Acad.

Hung., 69, 311–317.

71. Triebe, R.W. & Tezel, F. H. (1995).Adsorption of nitrogen, carbon

monoxide,carbon dioxide and nitric oxide on molecular sieves. Gas. Sep. Purif.

Vol. 9, No. 4, 223-230.

72. Vaart, R. V., Huiskes, C., Bosch, H. & Reith, T. (2000). Single and mixed gas

adsorption equilibria of carbon dioxide/methane on activated carbon. Adsorption,

6, 311–323.

73. Wang, Y. & LeVan, M. D. (2009). Adsorption Equilibrium of Carbon Dioxide

and Water Vapor on Zeolites 5A and 13X and Silica Gel Pure Components. J.

Chem. Eng. Data, 54, 2839–2844.

74. Yang, R. T. (2003). Adsorbents:Fundamentals and Applications. New Jersey,

John Wiley & Sons, Inc.

75. Younger, A.H. (2004). Natural Gas Processing Principles and. Technology - Part

I. Canada: University of Calgary.

76. Yucel, H. & Ruthven, D. M. (1980). Diffusion of CO2 in 4A and 5A Zeolite

Crystals. Journal of Colloid and Interface Science, 74, No. 1.

77. Zhang, J., Xiao,P., Li, G. & Webley, P. A. (2009). Effect of Flue Gas Impurities

on CO2 Capture Performance from Flue Gas at Coal-fired Power Stations by

Vacuum Swing Adsorption. Energy Procedia, 1, 1115–1122.

99
78. Zhang, Z., Zhang, W., Chen, X., Xia, Q. & Li, Z (2010). Adsorption of CO2 on

Zeolite 13X and Activated Carbon with Higher Surface Area. Separation Science

and Technology, 45, 710–719.

100
APPENDIX A

Experimental results of pure CO2 adsorption equilibrium


Table A.1 CO2 adsorption equilibrium data on zeolite 13X at different pressures and temperatures

T = 293 K T = 303 K T = 313 K T = 323 K T = 333 K


Adsorption Amount Adsorption Amount Adsorption Amount Adsorption Amount Adsorption Amount
pressure adsorbed pressure adsorbed pressure adsorbed pressure adsorbed pressure adsorbed
(atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g)
0.06 0.54 0.29 1.13 0.35 1.22 0.37 0.97 0.36 0.92
0.06 1.05 0.34 1.98 0.44 2.13 0.47 1.86 0.48 1.81
0.07 1.61 0.44 2.76 0.66 2.98 0.66 2.71 0.74 2.74
0.09 2.17 0.57 3.51 0.90 3.75 0.87 3.35 1.23 3.51
0.12 2.63 0.80 4.19 1.59 4.28 1.60 3.96 1.90 3.91
0.13 3.13 1.34 4.71 2.77 4.66 3.33 4.50 3.32 4.38
0.19 3.65 2.53 5.17 4.28 4.91 5.70 4.81 6.30 4.76
0.30 4.11 5.24 5.55 5.74 5.07 10.06 5.05 12.86 5.05
0.45 4.50 9.35 5.80 11.00 5.32 15.28 5.18 20.64 5.12
0.63 4.78 14.75 5.94 18.60 5.43 23.88 5.22 29.40 5.15
0.83 5.04 22.72 6.00 23.94 5.49 31.10 5.25 34.19 5.15
1.27 5.31 30.67 6.01 31.23 5.49 35.00 5.26
2.21 5.61 34.48 6.06 34.12 5.52
3.84 5.89
6.61 6.14
10.55 6.35
18.16 6.53
25.02 6.65
31.95 6.78
33.79 6.89
35.41 7.02

101
Table A.2 CO2 adsorption equilibrium data on zeolite 5A at different pressures and temperatures

T = 293 K T = 303 K T = 313 K T = 333 K


Adsorption Amount Adsorption Amount Adsorption Amount Adsorption Amount
pressure adsorbed pressure adsorbed pressure adsorbed pressure adsorbed
(atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g)
0.11 0.68 0.47 1.13 0.47 0.74 0.48 1.17
0.14 1.64 0.71 2.26 0.76 2.14 0.93 2.28
0.23 2.55 1.44 3.24 1.73 3.23 1.77 2.85
0.58 3.48 2.32 3.66 3.45 3.61 3.72 3.16
1.25 3.85 3.96 3.85 6.12 3.78 6.55 3.33
3.96 4.15 6.63 3.98 16.74 3.94 11.22 3.44
8.23 4.33 10.22 4.09 25.06 3.99 16.57 3.49
15.39 4.51 15.69 4.18 32.65 4.02 22.60 3.50
22.67 4.56 24.27 4.21 35.49 4.03 28.35 3.52
28.82 4.60 32.80 4.28 34.63 3.54
34.59 4.68 34.41 4.30

102
Table A.3 CO2 adsorption equilibrium data on zeolite 4A at different pressures and temperatures

T = 293 K T = 303 K T = 323 K T = 333 K


Adsorption Amount Adsorption Amount Adsorption Amount Adsorption Amount
pressure adsorbed pressure adsorbed pressure adsorbed pressure adsorbed
(atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g)

0.21 1.00 0.62 0.91 0.25 0.21 0.25 0.14


0.30 1.47 1.62 2.16 0.35 0.30 0.35 0.19
0.53 1.99 2.87 2.62 0.65 0.55 1.25 0.69
1.06 2.48 4.38 2.84 0.95 0.80 2.25 1.24
3.59 3.01 6.83 3.03 1.25 1.05 3.09 1.64
5.82 3.24 10.46 3.18 1.75 1.45 5.89 2.53
12.77 3.47 23.65 3.41 2.25 1.81 12.69 2.75
24.10 3.59 32.02 3.49 2.75 2.12 24.05 2.90
31.43 3.71 35.43 3.55 3.05 2.25 30.49 2.95
34.51 3.82 5.82 2.75 34.73 3.01
12.78 2.97
24.15 3.11
30.57 3.15
34.69 3.21

103
Table A.4 CO2 adsorption equilibrium data on MSC-3R at different pressures and temperatures

T = 293 K T = 303 K T = 313 K T = 323 K T = 333 K


Adsorption Amount Adsorption Amount Adsorption Amount Adsorption Amount Adsorption Amount
pressure adsorbed pressure adsorbed pressure adsorbed pressure adsorbed pressure adsorbed
(atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g)
0.34 0.63 0.60 0.74 0.84 0.95 0.69 0.54 1.07 0.75
0.54 1.10 1.11 1.50 1.38 1.47 1.48 1.29 2.31 1.58
0.75 1.57 1.82 2.04 2.30 2.03 2.74 1.94 4.89 2.23
0.95 1.93 2.78 2.50 4.56 2.65 4.37 2.40 8.12 2.60
1.25 2.23 4.60 2.96 7.97 3.05 6.63 2.74 12.59 2.89
2.11 2.79 6.58 3.22 10.61 3.21 10.46 3.02 21.24 3.04
4.92 3.43 9.64 3.44 15.42 3.34 14.24 3.15 30.50 3.08
7.98 3.73 11.88 3.54 18.86 3.39 19.07 3.21 34.78 3.11
12.60 3.93 16.66 3.63 23.67 3.40 24.31 3.22
18.62 4.06 22.85 3.63 27.01 3.40 28.14 3.23
27.27 4.11 30.48 3.64 31.64 3.40 32.39 3.23
33.72 4.18 34.53 3.66 34.48 3.41 35.03 3.24
35.36 4.22

104
Table A.5 CO2 adsorption equilibrium data on GCA-830 at different pressures and temperatures

T = 293 K T = 303 K T = 313 K T = 323 K T = 333 K


Adsorption Amount Adsorption Amount Adsorption Amount Adsorption Amount Adsorption Amount
pressure adsorbed pressure adsorbed pressure adsorbed pressure adsorbed pressure adsorbed
(atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g)
0.40 1.14 1.01 1.66 0.93 1.17 1.20 1.46 1.57 1.48
0.74 2.16 1.79 2.83 1.81 2.35 3.20 3.03 3.42 2.70
1.05 3.03 3.55 4.38 3.30 3.56 4.72 3.86 6.65 4.05
1.83 4.17 5.68 5.56 4.61 4.33 7.56 4.86 8.91 4.68
2.91 5.35 8.59 6.47 7.31 5.40 11.34 5.70 12.35 5.36
4.42 6.45 10.37 6.82 10.67 6.21 15.72 6.31 15.79 5.83
6.76 7.49 15.27 7.49 14.94 6.85 24.31 6.83 27.36 6.44
10.69 8.47 23.75 8.00 19.86 7.29 32.07 7.03 34.35 6.56
15.89 9.08 32.85 8.21 27.27 7.62 34.51 7.09
23.98 9.50 35.56 8.27 33.68 7.77
32.05 9.77 35.37 7.81
35.02 9.88

105
Table A.6 CO2 adsorption equilibrium data on GCA-1240 at different pressures and temperatures

T = 293 K T = 303 K T = 313 K T = 323 K T = 333 K


Adsorption Amount Adsorption Amount Adsorption Amount Adsorption Amount Adsorption Amount
pressure adsorbed pressure adsorbed pressure adsorbed pressure adsorbed pressure adsorbed
(atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g) (atm) (mmol/g)
0.40 0.87 0.60 1.03 0.70 1.08 0.77 0.91 1.08 1.06
0.65 1.72 1.02 1.86 1.19 1.87 1.47 1.73 1.88 1.82
0.85 2.55 1.76 3.02 2.02 2.80 2.65 2.69 4.80 3.50
1.16 3.28 2.65 3.90 3.45 3.92 4.75 3.85 7.44 4.42
1.95 4.52 3.81 4.80 6.35 5.33 8.77 5.13 10.65 5.17
3.45 6.05 4.98 5.52 10.12 6.38 16.32 6.26 13.23 5.63
4.77 6.90 6.62 6.23 13.78 7.00 24.42 6.71 20.46 6.35
7.88 8.09 9.64 7.10 20.12 7.58 33.31 6.92 31.30 6.67
11.95 8.96 13.83 7.78 25.74 7.84 35.37 6.77
16.32 9.47 19.62 8.28 32.67 7.99
23.20 9.87 25.30 8.54 34.44 8.02
30.33 10.07 31.49 8.64
34.07 10.25 34.76 8.69

106

You might also like