You are on page 1of 127

DESIGN METHOD FOR LAYERED BED ADSORBER FOR SEPARATION OF

CO2 AND N2 FROM NATURAL GAS USING ZEOLITE13X, CARBON

MOLECULAR SIEVE AND ACTIVATED CARBON

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

By

Mohammad Rokanuzzaman

Regina, Saskatchewan

February, 2015

Copyright 2015: Mohammad Rokanuzzaman


UNIVERSITY OF REGINA

FACULTY OF GRADUATE STUDIES AND RESEARCH

SUPERVISORY AND EXAMINING COMMITTEE

Mohammad Rokanuzzaman, candidate for the degree of Master of Applied Science in


Process Systems Engineering, has presented a thesis titled, Design Method for Layered
Bed Adsorber for Separation of CO2 and N2 from Natural Gas Using ZEOLITE13X,
Carbon Molecular Sieve and Activated Carbon, in an oral examination held on
December 16, 2014. 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. Daoyong Yang, Petroleum Systems Engineering

Supervisor: Dr. Amornvadee Veawab, Process Systems Engineering

Committee Member: Dr. Stephanie Young, Process Systems Engineering

Committee Member: Dr. Adisorn Aroonwilas , Industrial Systems Engineering

Chair of Defense: Dr. Doug Durst, Faculty of Social Work


ABSTRACT

Natural gas (NG) is a low-carbon fossil fuel that carries impurities such as carbon

dioxide (CO2) and nitrogen (N2). These two impurities reduce the heating value of NG.

Also, CO2 causes corrosion in the pipeline and N2 produces nitrogen oxide (NOx) when

combusted. These facts have forced NG transmission and distribution companies to limit

the concentrations (mole percent) of CO2 (≤ 3%) and N2 (≤4%). Consequently, selective

separation of CO2 and N2 from NG has gained considerable importance.

There are many technologies that are in use for separation of these two

constituents. Most of them are suitable for single component separation: either CO2 or

N2. In the context of multicomponent separation common in industries, adsorption is an

emerging technology that offers low-cost and energy-efficient separation for small- to

medium-sized industries. The technology lacks commercial availability due to its

dependency of design methodologies on experimentation, simulation or both.

This work focuses on easy-to-use design methodology for the design of a double

bed adsorber. This easy-to-use methodology is tailored for separation of CO2 and N2 from

NG using zeolite13X and a carbon molecular sieve (CMS3K) or activated carbon (ACB).

These adsorbents are commercially available, and they offer easy and energy efficient

regeneration for repeated uses. The product will meet the specified concentration limit for

NG transmission and distribution systems.

To achieve this goal, two layered bed adsorbers, zeolite13X-CMS3K and

zeolite13X-ACB, were simulated using Aspen Adsorption. Simulation requires a

trustworthy mathematical framework i.e. model. Therefore, a model was developed in

i
Aspen adsorption by selecting relevant equations and submodels. Inputs for the model

were collected from literature, calculated using various equations, and obtained by fitting

experimental data. A numerical solution method was specified and, finally, the model

was validated against experimental measurements.

A parametric study was performed for a wide range of operating conditions. Data

generated through parametric study were correlated. The correlations, the first of this

kind, can be used to predict required amounts of adsorbents for 100% CO2 separation and

50 to 90% N2 separation. Finally, a procedure was outlined to transform the amount of

adsorbent into the physical dimensions of an adsorber.

ii
ACKNOWLEDGEMENTS

I would like to express my sincere gratitude to my supervisor, Dr. Amornvadee

Veawab, for giving me the opportunity to carry out this interesting research under her

enthusiastic supervision. Her enormous financial and technical support, valuable

guidance, and encouragement were a great source of inspiration and the driving force

throughout the entire course of this research. I would also like to express my gratitude to

Dr. Adisorn Aroonwilas for his great support, guidance, and encouragement.

I would like to thank the Natural Sciences and Engineering Research Council of

Canada (NSERC), SaskEnergy, and the Faculty of Graduate Studies and Research

(FGSR) for their financial support. I would also like to thank the Faculty of Engineering

and Applied Science at the University of Regina for their help and support.

Finally, I am sincerely thankful and grateful to my parents, family, and friends

for their unconditional love, prayers, and support to fulfill my dreams.

iii
Table of contents

Abstract i

Acknowledgement iii

Table of contents iv

List of Tables vii

List of Figures ix

Nomenclature xii

1. Introduction 1

1.1 Natural gas 1

1.2 Industrial separation processes for removal of N2 and CO2 from natural gas 3

1.3 Adsorption process and adsorbents for CO2 and N2 removal 5

1.4 Modeling and simulation 7

1.5 Research motivation, objectives and scope of work 9

2. Literature review 12

2.1 Scope of review 12

2.2 Adsorption fundamentals 12

2.3 Adsorbents 13

2.3.1 Zeolite13X 14

2.3.2 Carbon adsorbents 15

2.4 Adsorption modeling 17

2.5 Multicomponent separation 20

2.6 Numerical Solution of partial differential equations 22

3. Modeling and simulations of a gas adsorber 25

iv
3.1 Adsorption modeling 25

3.1.1 Model equations 27

3.1.2 Solution of model equations 34

3.1.3 Calculation procedure 35

3.2 Model validation 38

3.2.1 Nitrogen separation using activated carbon 38

3.2.2 Methane separation from hydrogen using Zeolite 5A 42

3.2.3 Carbon dioxide separation using Zeolite 13X 45

4. Results and Discussion 50

4.1 Description of simulated gas adsorption systems 50

4.2 Simulation results for zeolite13X 52

4.2.1 Parametric study 56

4.2.2 Correlation to determine amount of zeolite13X 61

4.3 Simulation results for zeolite13X-CMS3K system 64

4.3.1 Parametric Study 68

4.3.2 Correlations based on simulated results 73

4.4 Simulation results for zeolite13X-ACB system 77

4.4.1 Parametric study 81

4.4.2 Correlations based on simulated results 86

4.5 Determination of column dimensions using correlations 89

5. Conclusions and recommendation for future work 92

5.1 Conclusions 92

5.2 Recommendation for future work 94

v
References 95

Appendix – A: Adjustments of transport parameters 106

vi
List of Tables

Table 1.1 Quantity of air pollutants produced from fossil fuel combustion in 2

lbs/billion Btu (U.S. Energy Information Administration (EIA),

1999)

Table 1.2 Composition of natural gas observed in different reservoirs as 2

mole percentage (Kidnay and Parish, 2006)

Table 1.3 Specification of pipeline natural gas (modified from Kidnay and 4

Parish, 2006)

Table 3.1 Model equations 26

Table 3.2 Model input (N2-ACB system) 40

Table 3.3 Model inputs (CH4-H2-zeolite5A system) 43

Table 3.4 Model inputs (CH4-CO2-N2-zeolite13X system) 46

Table 3.5 Isotherm parameters (CH4-CO2-N2-zeolite13X system) 47

Table 4.1 Physical properties of zeolite13X and properties of column 53

(Cavenati et al., 2006)

Table 4.2 Parameters used in simulation for zeolite13X 54

Table 4.3 Parameters of correlation for determination of amounts of 62

zeolite13X

Table 4.4 Physical properties of double bed adsorber (zeolite13X-CMS3K) 65

(Cavenati et al., 2006)

Table 4.5 Parameters used in simulation of zeolite13X-CMS3K system 66

Table 4.6 Parameters for correlations 4.2 75

vii
Table 4.7 Physical properties of double bed adsorber (zeolite13X-ACB) 78

(Cavenati et al., 2006 and Shen et al., 2010)

Table 4.8 Parameters used in simulation of zeolite13X-ACB system 79

Table 4.9 Parameters for correlation 4.3 87

viii
List of Figures

Figure 3.1 Calculation procedure 37

Figure 3.2 Breakthrough concentration profiles of N2 in pitch-based AC 41

beads (0.5% N2 in helium at 303K and 1 bar) under isothermal

conditions

Figure 3.3 Breakthrough concentration profile of methane in Zeolite 5A 44

(8.8% Methane in Hydrogen at 303K and 20.2 bar) under

isothermal conditions

Figure 3.4 Breakthrough concentration profiles of 70% CH4, 20% CO2 and 48

10% N2 in Zeolite 13X at 300K and 2.5 bars

Figure 3.5 Temperature profile at bed exit (70% CH4, 20% CO2 and 10% 49

N2 in Zeolite 13X at 300K and 2.5 bars)

Figure 4.1 Double bed adsorber 51

Figure 4.2 Required amount of zeolite13X for complete separation of CO2 55

as a function of feed pressure

Figure 4.3 Adsorption capacities and selectivity for CO2-N2-zeolite13X 57

system

Figure 4.4 Effect of concentration (%) of CO2 on required amount of 59

zeolite13X for 100% separation of CO2

Figure 4.5 N2 separation efficiency of zeolite13X 60

Figure 4.6 Comparison of simulated and predicted amounts of zeolite13X 63

Figure 4.7 Total amount (kg/mol of feed gas) of adsorbents for N2 and CO2 67

ix
separation from natural gas for zeolite13X-CMS3K adsorber

Figure 4.8 Effect of feed pressure on N2 separation efficiency (%) for 68

zeolite13X-CMS3K system

Figure 4.9 Effect of feed gas pressure on total amount of adsorbents for 70 70

to 90% N2 separation efficiency for zeolite13X-CMS3K system

Figure 4.10 Effect of feed concentration on nitrogen separation efficiency at 72

2.5 bars for zeolite13X-CMS3K system

Figure 4.11 Effect of N2 separation efficiency on total amount of adsorbent 74

for zeolite13X-CMS3K system

Figure 4.12 Comparison of simulated result with the results obtained from 76

correlation 4.2 for feed composition of 75% CH4, 15% N2 and

10 % CO2 for zeolite13X-CMS3K system

Figure 4.13 Total amount of adsorbents for N2 separation at different feed 80

pressures and compositions (zeolite13X-ACB system)

Figure 4.14 Effect of feed pressure on total amount of adsorbent for different 82

N2 separation efficiency (zeolite13X-ACB system)

Figure 4.15 Effect of concentration on total amount of adsorbents at different 84

feed pressures for zeolite13X-ACB system

Figure 4.16 Effect of N2 separation efficiency on total amount of adsorbent 85

for feed pressures of (a) 2.5 bars and (b) 30 bars (zeolite13X-

ACB system)

Figure 4.17 Comparison of simulated and predicted (correlation 4.3) results 88

Figure 4.18 Calculation procedure for determination of column dimension 91

x
using correlations

Figure A.1 Breakthrough of CO2 in zeolite13X for various mass transfer 107

resistances

Figure A.2 Breakthrough of CO2 in zeolite13X with modified macropore 109

resistance

Figure A.3 Effect of conductivity (gas and solid) on breakthrough dynamics 111

xi
Nomenclature

C0 Initial concentration (kmol/m3)

Ci Molar concentration (kmol/m3)

Cpa Specific heat capacity of adsorbed phase (MJ/kmol/K)

Cps Specific heat capacity of adsorbent (MJ/kmol/K)

Cpw Specific heat capacity of column wall (MJ/kg/K)

Cvg Specific gas phase heat capacity at constant volume (MJ/kmol/K)

Da Axial dispersion coefficient (m2/s)

DAB Binary diffusivity (cm2/sec)

Db Bed diameter (m)

Dk Knudsen diffusivity

Dm Molecular diffusivity

Dp Pore diffusivity (m2/s)

Hamb Wall-ambient heat transfer coefficient (MW/m2/K)

Hw Gas-wall heat transfer coefficient (MW/m2/K)

K Dimensionless Henry’s constant

KH Henry’s constant

M Molecular weight (kg/kmol)

Nu Nusselt number

P Feed pressure (bar)

PeH Peclet number for gas wall heat transfer

Pr Prandtl number

xii
Q Amount of adsorbents (g or kg per mol of feed gas)

Rc Crystal radius (m)

Re Reynolds Number

S Separation factor (mol/mol)

Sc Schmidt Number

Sh Sherwood Number

T0 Wall temperature (K)

Tamb Ambient temperature (K)

Tg Gas phase temperature (K)

Ts Solid phase temperature (K)

Tw Wall temperature (K)

V Atomic diffusion volume

WT Width of column (m),

Z Height of adsorbent bed (m)

ap Specific particle surface per unit volume bed (m2 (Particle area)/m3 (Bed))

dp particle diameter (m)

hf Gas-solid heat transfer coefficient (MW/m2/K)

k Effective mass transfer coefficient

kf Film mass transfer coefficient (m/s),

kg Conductivity of gas phase

q0 Initial loading (kmol/kg)

q Loading (kmol/kg)

q* Instantaneous equilibrium concentration (kmol/kg)

xiii
rmac Macropore radius

rp Particle radius (m)

t Time (second)

w Solid phase concentration (kmol/kg)

yi Mole fraction of component i

εb Bed voidage

εi Interparticle voidage

νg Gas velocity (m/s)

μ Dynamic viscosity (N s/m2)

η Separation efficiency (%)

μmix Viscosity of gas mixture

ψ Shape factor

ρg Gas phase molar density (kmol/m3)

ρmix Density of gas mixture

ρs Adsorbent bulk density (kg/m3)

ρw Wall density (kg/m3)

φij Binary viscosity of gas mixture

τ Tortuosity

ΔHi Heat of adsorption (MJ/kmol).

xiv
1 Introduction

1.1 Natural gas

More than 80% of energy comes from carbon-based fossil fuel (coal, oil, and

natural gas), and such contribution is expected to remain the same until 2030 (World

Energy Council, 2010). Combustion of these fossil fuels produces carbon dioxide (CO2),

which has already raised concern regarding global warming and climate change. Some

other pollutants that are also associated with fossil fuel combustion are carbon monoxide

(CO), nitrogen oxides (NOx), sulfur dioxide (SO2), particulate matters (PMs),

formaldehyde (CH2O), and mercury (Hg) (Table 1.1). These pollutants pose adverse

impacts on human health and the environment. Among the fossil fuels, natural gas is

considered to be the cleanest fossil fuel as it produces fewer quantities of CO2, NOx, SO2,

PMs, and Hg than coal and oil (EIA, 1999).

Natural gas is a complex mixture of hydrocarbons (such as methane, ethane,

propane, butane, and heavier hydrocarbons) and nonhydrocarbons (such as N2, CO2, and

hydrogen sulfide (H2S)). It may be present as free gas (bubbles) or dissolved in either

crude oil or brine under reservoir conditions in hundreds of different components with

various concentrations. Even two wells in the same reservoir may yield different natural

gas compositions (Younger, 2004). For example, as shown in Table 1.2, the

concentrations of CH4 vary from 29.98 to 96.91%, while the concentrations of N2 and

CO2 vary from 0.68 to 26.1% and 0.82 to 42.66%, respectively. Some reservoirs other

than those shown in Table 1.2 may have extreme contents of CO2 (92%), N2 (86%), and

H2S (88%) (Hobson and Tiratso, 1985).

1
Table 1.1: Quantity of air pollutants produced from fossil fuel combustion in lbs/billion

Btu (U.S. Energy Information Administration (EIA), 1999)

Pollutant Natural Gas Oil Coal


Carbon dioxide 117000 164000 208000
Carbon monoxide 40 33 208
Nitrogen oxide 92 448 457
Sulfur dioxide 0.60 1122 2591
Particulates 7 84 2744
Formaldehyde 0.75 0.22 0.221
Mercury 0.0005 0.007 0.016

Table 1.2: Composition of natural gas observed in different reservoirs as mole percentage

(Kidnay and Parish, 2006)

Rio
Bach
Component Canada Western Southwest Miskar Cliffside Arriba
Ho
(Alberta) Colorado Kansas Tunisia Texas New
Vietnam
Mexico
Methane 77.1 29.98 72.89 70.85 63.90 65.80 96.91
Ethane 6.60 0.55 6.27 13.41 3.35 3.80 1.33
Propane 3.10 0.28 3.74 7.50 0.96 1.70 0.19
Butane 2.00 0.21 1.38 4.20 0.54 0.80 0.05
Pentane and 3.00 0.25 0.62 2.64 0.63 0.50 0.02
heavier
Helium 0 0 0.45 0 0 1.8 0
Nitrogen 3.20 26.10 14.65 0.21 16.90 25.60 0.68
Carbon dioxide 1.70 42.66 0 0.06 13.58 0 0.82
Hydrogen sulfide 3.30 0 0 0 0.09 0 0

2
Natural gas is typically processed to meet pipeline specifications (Table 1.3),

which are intended to deliver the natural gas with high heating value to the end users and

also to protect pipeline from corrosion and plugging. For example, to prevent corrosion,

the concentrations of CO2, H2S, and mercaptans or total sulfur are limited to less than 3%

(mole), 6–7 mg/m3, and 115–460 mg/m3, respectively, while to prevent liquid dropout,

the concentrations of butane and heavier hydrocarbons are limited to less than 2.0%

(mole) and less than 0.5%, respectively. This study focuses on separation of CO2 and N2

from natural gas for the purpose of compliance to the pipeline gas specification. These

two gases are considered to be the contaminants of natural gas since they have no heating

value and occupy transport volume. The CO2 corrodes pipelines in the presence of water

and the N2 produces NOx when natural gas is combusted.

1.2 Industrial separation processes for removal of N2 and CO2 from natural gas

Four gas separation techniques, namely, cryogenic distillation, membrane

separation, absorption, and adsorption are in practice for natural gas purification. Of

these, the cryogenic and absorption processes are economically viable at high gas

throughputs (> 15 MMscfd), while the membrane and adsorption processes are viable at a

gas throughputs of 0.5 - 25 MMscfd and 2 - 15 MMscfd, respectively (Kidnay and Parish,

2006). The absorption process is widely used for CO2 capture while the cryogenic

process is established for N2 removal from natural gas. Neither of these two processes is

suitable for removal of both CO2 and N2 simultaneously. The cryogenic process for N2

separation requires extensive pretreatments that eliminate CO2 from feed gas.

3
Table 1.3: Specification of pipeline natural gas (modified from Kidnay and Parish, 2006)

Components Quantity (mole % or as mentioned)


Methane 75.0% (minimum)
Ethane 10.0% (maximum)
Propane 5.0% (maximum)
Butane 2.0% (maximum)
Pentanes and heavier 0.5% (maximum)
Nitrogen 4.0% (maximum)
Carbon dioxide 3.0% (maximum)
Hydrogen sulfide 6 to 7 mg/m3
Total sulfur 115-460 mg/m3
Water vapor 60-110 mg/m3

4
Additionally, these two processes incur high operating costs compared to membrane or

adsorption processes (Robertson, 2007).

Conventional membrane (cellulose acetate/polysulfone) separation technologies

use kinetic diameters of molecules as the separation criterion. The kinetic diameters of

CH4, CO2, and N2 are 3.8Å, 3.3Å, and 3.6Å, respectively (Do, 1998). These diameters are

too close to offer a favorable selectivity for membrane. Silicone membrane separation

uses equilibrium affinity as the separation criterion. In CH4-N2 separation using this

membrane, CH4 comes out at low pressure end and, hence, leads to additional

recompression costs. Another critical element of the process is the pretreatment of the

feed gas since particulates block the membrane openings and liquids cause swelling,

resulting in decreased performance and even physical damage. Membranes can be highly

efficient mass-separating mediums, especially when the species that are to pass through

the membrane are present in a large concentration (Choi et al., 2009).

Adsorptive separation is a process where certain fluid particles are bonded to the

surface of an adsorbent by physical/chemical bonding. It is based on three distinct

mechanisms: steric (dimension: pore and molecule size), equilibrium (accommodation

ability), and kinetic (diffusion rate) mechanisms (Do, 1998). The first step in separation is

adsorption during which species are preferentially picked up from the feed by

adsorbent/adsorbents (porous solid), and the second is regeneration or desorption during

which the species are removed from the adsorbent. There are two types of adsorption

processes: physical adsorption and chemical adsorption. Of them, the physical-adsorption

process is an energy efficient and low cost technology (Siriwardane et al., 2001).

5
1.3 Adsorption process and adsorbents for CO2 and N2 removal

The separation efficiency of the adsorption process depends on the quality of the

adsorbent, a porous solid. Ideally, an adsorbent should have large adsorption capacity,

fast adsorption and desorption kinetics, infinite regenerability, and a wide yet tunable

range of operating conditions (Choi et al., 2009). However, in practice, it is rare to find

such an ideal adsorbent. Another adsorption behavior to consider is the competitive

adsorptions, known as selectivity, of components (CO2/N2/CH4) of a gas mixture (natural

gas, considered to be a mixture of CH4, CO2, and N2 for simplicity). Thus, optimizing the

trade-off between beneficial and non-beneficial features is the key in process design and

operation.

The adsorbents that have been used for CO2 separation are zeolites (crystalline

aluminosilicates), activated carbons, calcium oxides, hydrotalcites, and supported amines.

A review of these materials can be found in Choi et al. (2009). The zeolite-based

adsorbents were reported to yield relatively high adsorption capacities (Ding and Alpay,

2000). Harlick and Tezel (2004) carried out an experimental screening study of various

synthetic zeolite adsorbents and reported that zeolite13X possesses a maximum CO2

adsorption capacity of 4.5 mol/kg at 1 bar and 295K. Typically, the zeolites recover fresh

adsorption capacity when regenerated, though little irreversible behavior was reported by

Brandani and Ruthven (2004). Zeolite13X also provides high selectivity for CO2 over

CH4 and N2 (Cavenati et al., 2004).

The adsorption of N2 on several commercial adsorbents was studied by many

researchers. Of these, carbon-based adsorbents, such as the carbon molecular sieve

(CMS) and activated carbon bead (ACB), showed greater adsorption capacity (0.27

6
mol/kg at 303K and 100 kPa) (Shen et al., 2010). The notable difference between these

two adsorbents is in pore distribution. ACB carries both micropores and transitional pores

ranging from 10 to 500 Angstroms (Å), while CMS contains uniform pores of less than

10 Å (Do, 1998). The porous structures of ACB and CMS lead to equilibrium-based

separation and kinetic-based separation, respectively. In equilibrium-based separation,

the equilibrium selectivity of these carbonaceous adsorbents favors CH4 over N2, which

eventually renders more adsorbed CH4 at the surface of the adsorbents. In kinetic-based

separation, the kinetic selectivity of N2 over CH4, offered by CMS, leads to less

adsorption of CH4 in adsorbents. An advantage of equilibrium-based separation is that it

offers longer cycle time than its counterpart: kinetic separation. Both ACB and CMS

have greater affinity for CO2 than either CH4 or N2, which leads to the necessary removal

of CO2 from the natural gas to facilitate optimum nitrogen separation. In this study,

zeolite13X was used to separate CO2 from a ternary mixture of CH4, CO2, and N2. ACB

and CMS were used to separate N2 from a binary mixture of CH4 and N2.

1.4 Modeling and simulation

In adsorption separation systems, the process variables are strongly coupled,

resulting in complex interrelationships. Therefore, the effect of any single variable on

separation efficiency is simply unpredictable by simple reasoning or empiricism (Hassan

et al., 1986). Furthermore, a change in adsorbent material adds additional complexity due

to their unique adsorbate-adsorbent behaviour under the same operating conditions

(Flores-Fernandez and Kenney, 1983). Thus, the design and optimization requires either

7
extensive experimentation or the guidance of a predictive model (Farooq and Ruthven,

1991).

An adsorption model requires in-depth mechanistic knowledge of the kinetics and

equilibria of adsorption process and their impact on the dynamic response of an

adsorption column (Ruthven, 2000). The pioneer studies of kinetics and equilibria

include, but are not limited to, the works of Habgood (1958), Barrer et al. (1963), and

Mayers and Prausnitz (1965). These studies reveal that the adsorption separation

efficiency is controlled by either equilibrium or kinetics.

The simplest adsorption model uses the equilibrium theory. Thomas (1944) can be

credited to be the pioneer of the use of equilibrium theory in an ion exchange column.

His work was later shaped by Glueckauf (1955) and Rosen (1952) in a general form for

application in gas adsorption processes. Such an equilibrium model accommodates the

analytical solution of the governing material balance equations and provides useful

behavioral insights. However, the theory does not consider real situations such as partial

equilibrium and dispersive flows observed in an industrial setup. This model also ignores

the mass transfer resistances (Hassan et al., 1986). Failure of incorporating such

important process characteristics resulted in outsized deviations in the adsorption of CO2

on silica gel (Mitchell and Shendahnan, 1972) and that of ethylene on zeolite 4A/5A

(Hassan et al., 1985).

The limited success of equilibrium models necessitates consideration of kinetic

models as well. A kinetic model requires adequate representation of mass transfer

kinetics. Mitchell and Shendahnan (1973) adopted this approach with one mass transfer

resistance and constant velocity. They laid the foundation of a dynamic model that, later

8
on, comprehensibly described the mass transfer kinetics as well as resistances (Hassan et

al., 1986). This dynamic model accounts for realistic scenarios, such as axial mixing and

mass transfer resistances, which are always likely to be present in the practical systems.

The model is, therefore, more realistic and sufficiently general to be applied for detailed

optimization studies of both systems.

The dynamic model equations then become very complicated and, hence, were

solved numerically. Various numerical solution procedures were facilitated by the use of

computers in the early 1980s. This trend gradually paved the way for the development of

process simulators (Ruthven, 2000). The research performed by Liapis and Crosser

(1982), one of the earliest examples, served as the foundation of commercial simulators

such as Aspen Adsorption (Nilchan and Pantelides, 1998). The availability of such

simulators made it possible to simulate the adsorption process with more rigorous

mathematical models by greatly reducing the burden of the manual handlings of complex

equations and their numerical solutions. The use of such simulators is not limited to

merely solving some equations but rather has expanded into the design and optimization

of commercial processes.

1.5 Research motivation, objectives and scope of work

The separation of CO2 and N2 from NG is a two-step separation process as two

adsorbents, CO2 selective and N2 selective, are required. This can be done in a single

column using layers of different adsorbents (Chlendi et al., 1995; Chlendi and Tondeur,

1995; Malek and Farooq, 1998; Yang and Lee, 1998; Lee et al., 1999; Jee et al., 2001;

Takamura et al., 2001; Cavenati et al., 2006; Rebeiro et al., 2008) or columns in series

9
carrying different adsorbents (Sircar, 1979; Kumar, 1990). Of these two types of adsorber

combination, the layered bed adsorber offers compact design and operation flexibility.

Layered bed adsorption systems were studied by many researchers, as mentioned

above, for the separation of various components, such as CO2, N2, CH4, CO, on various

adsorbents. A discussion on the layered bed adsorption system is included in Chapter 2.

Of the studies mentioned, the most relevant study for the separation CO2 and N2 from NG

was published by Cavenati et al. (2006). They studied a layered bed adsorber containing

zeolite13X and CMS3K in terms of product purity and separation efficiency and the

effects of the ratio of bed height on separation efficiency. It was concluded that the bed

ratio has an insignificant effect on product purity and separation efficiency. For their

study, they used a single feed pressure (2.5 bars) and two compositions (mole %) of feed

gas (70% CH4/20% CO2/10% N2 and 60% CH4/20% CO2/20% N2). No conclusions were

drawn for other concentrations or other feed pressures. This study is good for conveying

an understanding of the adsorption behavior of CO2 and N2 in a mixture of CH4-CO2-N2.

No methodology for the design of a double bed adsorber was outlined.

To this end, this study aims to develop an easy-to-use design methodology for a

layered bed adsorber using commercial adsorbents. The method shall cover a wide range

of operating conditions consistent with CO2 and N2 concentrations observed in various

reservoirs and typical feed pressure of NG distribution pipelines. Also, the product shall

meet the concentration limit of CO2 and N2 in an NG distribution network to maintain

pipeline integrity and product purity. The method shall significantly reduce the need for

extensive experiment and simulation.

10
To achieve such an objective, a standalone simulation study was performed on a

two-layered bed adsorption system (zeolite13X-CMS3K and zeolite13X-ACB) in Aspen

Adsorption, a commercial simulator (discussed in Chapter 2). Since such study requires a

trustworthy mathematical framework or model, a model was developed using the

resources of the simulator. Required inputs were gathered from the literature, calculated

using various equations/correlations, and obtained through fitting the experimental data.

To justify the reliability of the model, it was validated against several adsorption

processes that covered various operating conditions on different adsorbents.

Parametric studies were performed for a wide range of operating conditions that

covered concentrations observed in various NG reservoirs and feed pressures of typical

NG distribution networks. The data generated through parametric study were correlated.

The correlation, the first of this kind, predicts the amount of adsorbents for 100% CO2

separation and 50-90% N2 separation. A step-by-step procedure was outlined to

transform the amount of adsorbent into physical dimensions of the adsorption column.

11
2 Literature Review

2.1 Scope of review

Charles W. Skarstorm was awarded the first patent on a commercial adsorptive

separation process for air fractionation in 1960. Since then, the technology has gained a

phenomenal growth in commercial applications and process concepts (Sircar, 2006). This

chapter addresses process fundamentals and essential components of process design that

has led the technology to the present state.

2.2 Adsorption fundamentals

Adsorption is a surface phenomenon that refers to enrichment (or rise in density)

of material at the vicinity of fluid-solid interfaces through physical or chemical bonding

(Rouquerol et al., 1999). The fluid is referred as an adsorbate and the solid is referred to

(porous and permeable) as adsorbents. An adsorbent selectively adsorbs a component or

components from a mixed feed. Such selective adsorption may depend on the difference

in adsorption at equilibrium or on a difference in adsorption rates.

There are three distinct mechanisms through which adsorption separation takes

place (Yang, 2003). They are (i) steric or molecular sieving effects, (ii) kinetic or

diffusional effects, and (iii) equilibrium or selective adsorption effects. The steric effect

allows small and properly shaped molecules to diffuse into adsorbent where the

molecules are consequently adsorbed while other molecules are barred from entering the

pores. The success depends on the pore diameter of adsorbents and the kinetic diameter

and shape of fluid particles. Examples of steric separation include gas drying with 3A

12
zeolite and the separation of normal paraffins from iso-paraffins and other hydrocarbons

with 5A zeolite (Yang, 1987). Kinetic separation is achieved by virtue of differences in

diffusion rates and, hence, the mechanism is also known as partial molecular sieving

action. For effective separation, the pore size needed to be precisely controlled between

the kinetic diameters of the molecules to be separated (Yang, 2003). Nitrogen-methane

separation with 5A zeolite and nitrogen-oxygen separation with a carbon molecular sieve

are examples of kinetic separations. An equilibrium effect uses the adsorbate-adsorbent

interaction at the solid surface when all the components of a gas mixture are present. The

strength and affinity of fluid particles determines selective adsorption of components.

Separation of carbon dioxide and methane with zeolite is an example of equilibrium

separation. In a practical process, any of the mechanisms or any combination of these

mechanisms may play a significant role since all of the mechanisms depend on the

geometry and topology of the adsorbent (Rigby et al., 2004).

2.3 Adsorbents

The essential component of an adsorption separation process is the adsorbent. As

for industrial applications, it is a structured solid with inter-connected voids that hold a

certain fluid and, hence, separates a contaminant from the bulk of fluids. Characteristics

of adsorbents have been described by various researchers. A summary can be found

elsewhere (Rigby et al., 2004). The description includes the origin, size, structure, and

inter-connectivity of pores. The portrayal of pores in terms of size, distribution, and inter-

connectivity has found widespread applications in industries. An adsorbent with

interconnected pores of the same distribution is known as a homogeneous adsorbent

13
(silica gel, activated carbon, activated alumina, etc.). In contrast, heterogeneous (carbon

molecular sieve, zeolite), also known as composite, adsorbent consists of pellets of

microporous crystal that result in a bidispersity in pore networks.

Several features illustrate the quality or usefulness of an adsorbent. In general, an

ideal or hypothetical adsorbent should have large adsorption capacity, fast adsorption and

desorption kinetics, infinite regenerability and stability, and a wide yet tunable range of

operating conditions (Choi et al., 2009). In reality, no single ideal adsorbent is likely to

exist and an effective separation process uses trade-offs of these features. Three

adsorbents (zeolite13X, activated carbon, and a carbon molecular sieve) were considered

in this study. Relevant discussion on these three adsorbents is included in next three

subsections.

2.3.1 Zeolite13X

Zeolites are tridimensional aluminosilicates: a periodic array of SiO4/AlO4

composed of Si and Al tetrahedra linked through bridging oxygen atoms giving rise to a

periodic distribution of pores and cavities of particular molecular dimensions. This

microporous adsorbent represents a major breakthrough in the adsorption separation

process due to their uniform pore structure (8 to 10 Å), wide topology, and high (thermal,

hydrothermal, and chemical) stability. There are different criteria (pore aperture, shape of

pores, dimensionality of channel, and channel connection) according to which the

structure of zeolites can be classified. Zeolites can be found in nature or can be

synthesized. In synthesized zeolites, the pore structures are controlled by replacing

14
negatively charged alumina with cations. Such replacement by sodium produces zeolite-

13X with a pore diameter of 8Å.

Adsorption separation of CO2 in Zeolite-13X has been studied by many

researchers. These studies were focused on three important characteristics of the

adsorption process: (i) nature of adsorption (Ward and Habgood, 1966), (ii) equilibrium

adsorption capacity (Siriwardane et al., 2005; Cavenati et al., 2005), and (iii)

breakthrough behavior (Cavenati et al., 2006). As per the study by Ward and Habgood

(1966), the dominant adsorption process in zeolite13X is physisorption and, hence, it

offers fresh adsorption capacity (when regenerated) and lower regeneration cost. As per

Siriwardane et al. (2005), zeolite13X offers the highest equilibrium adsorption capacity

among the adsorbents the tested.

The separation of molecules by zeolites as adsorbents can take place because of a

molecular sieve effect or selective adsorption. Though zeolites are known for molecular

sieving actions, separation of CO2 from a ternary gas mixture of CH4-CO2-N2 occurs due

to selective adsorption since kinetic diameter of CH4 (3.8Å), CO2 (3.3Å), and N2 (3.6Å)

are considerably less than the pore opening (8Å) of Zeolite-13X. All these three gases

get adsorbed in zeolite13X showing the highest capacity and selectivity (CO2 over CH4

or CO2 over N2) for carbon dioxide (Cavenati et al., 2004).

2.3.2 Carbon adsorbents

Carbon adsorbents are employed to absorb non-polar or weakly polar organic

molecules. They are roughly divided into four categories: (i) activated carbons (ACs), (ii)

carbon molecular sieves (CMS), (iii) activated carbon fibers (ACFs), and (iv) carbon-

15
based nanomaterials such as single-walled carbon nanotubes (SWNTs) (Tagliabue et al.,

2009). Among them, ACs and CMSs are the most employed material in industrial gas

separations. Despite favorable properties, high costs of ACFs and SWNTs limit their uses

to small units. ACs and CMSs were studied for nitrogen separation from NG by, for

example, Shen et al. (2010) and Cavenati et al. (2006). They also included a comparison

with other adsorbents with respect to N2 rejection.

Activated Carbon (AC): AC is a form of carbon processed to be riddle with small,

low-volume pores that increase the surface area available for adsorption or chemical

reactions. Their usefulness is undoubtedly derived from large pore volume as well as high

surface area (Yang, 2003). These meso- or micro-porous carbonaceous materials offer

advantages over other materials in terms of cost (Choi et al., 2009). Among practical

adsorbents that are being used in industries, activated carbons are complex in terms of

both pore structure and surface chemistry due to presence of slit-shaped micropores (3 to

10Å) and oxygen (Do, 1998). The adsorption properties of activated carbon depend on

raw material and, also, on activation process (Choi et al., 2009) as well as adsorbate-

adsorbent interactions. AC performs adsorption separation by exploiting differences in

equilibrium adsorption for the constituent of a gas mixture.

Carbon Molecular Sieve (CMS): CMSs are nanoporous materials that separate

adsorbing molecules on the basis of their size and shape. A noteworthy feature of the

CMS materials is that they separate molecules on the basis of rates of adsorption (Foley,

1995). In terms of molecular sieving, CMSs are similar to zeolites with distinctive

physical structures. For example, CMSs are amorphous solids that has no long-range

16
order while zeolites are crystalline ordered material. Another feature that sets CMSs apart

from zeolites is its variable surface chemistry (acidic/basic/neutral/radical).

2.4 Adsorption modeling

Mathematical exploration of adsorption processes traces back to work of Thomas

(1944) who studied the mechanism of ion exchange with zeolite in an ion-exchange

column. His analytical solution assumed a single solute solution. The work was then

extended to binary and multicomponent systems by Glueckauf (1949) with an assumption

of local equilibrium. With same assumption i.e. local equilibrium, Lapidus and

Amundsen (1952) examined the effects of longitudinal diffusion and incorporated first

order kinetics in their solution. In the same year (1952), Rosen published a study that

included the exact same solution of an adsorption model that introduced rate of

adsorption. He assumed that the rate of adsorption is linear.

LDF Model: The linear rate of adsorption was then explored by Glueckauf (1955)

in the form of a linear driving force (LDF) model. The LDF approximation founded the

basis for the kinetic model. Glueckauf (1955) introduced a value of 15 for the LDF

constant that provided satisfactory solutions for processes with large cycle times. For

smaller cycle times, Nakao and Suzuki (1983) proposed a graphic correlation that

provides the values of the LDF coefficient as a function of dimensionless time. Haynes

and Sharma (1975) incorporated more realistic cases of mass transfer limitations such as

film resistance, interparticle resistance, and intraparticle resistance in LDF

approximation. Serbezov and Sotirchos (1996) formulated a general methodology for the

17
development of LDF approximations of different degrees of complexity for

multicomponent mixtures.

Particle-bed Model: The particle-bed models are the most complex models for

adsorption-based separations as they combine equations for both the bed and the particle

(Serbezov and Sotirchos, 1999). This coupled approach was first formulated by Yang and

Doong in 1985. They assumed parabolic intraparticle concentration profiles in the

solution scheme, which is equivalent to Glueckauf’s LDF approximation (Serbezov and

Sotirchos, 1999). The model equations were also solved by many others using different

numerical approaches such as orthogonal collocation (Shin and Knaebel, 1987; Lu et al.,

1992), finite element method (Kikkinides and Yang, 1993), finite difference method (Sun

et al., 1996), and global collocation (Khrisnan, 1993).

Mass Transport Model: The mass transport models used in the formulation of the

equations for the adsorbent bed and the adsorbent particles are essential parts of the

overall model. In general, there are four mechanisms of mass transport that have to be

considered: bulk diffusion, Knudsen diffusion, Knudsen flow, and viscous flow.

Serbezov and Sotirchos (1997a) showed that, in the adsorbent bed, the dominant mode of

mass transport is typically viscous flow, which can be modeled by Darcy’s law. In the

adsorbent particles, however, depending on the operating conditions, all four mechanisms

may be equally important (Serbezov and Sotirchos, 1997b). Therefore, the intraparticle

mass transport model must accurately describe the multicomponent mass transport of the

individual species over a wide range of conditions in order to be useful for simulations.

The widely applied and accepted model for the intraparticle mass transport in the

adsorption literature so far is the Fickian model, which provides a simple mathematical

18
expression for the molar fluxes of the species. However, the Fickian model does not

account for intraparticle viscous transport and underestimates the Knudsen flow of each

species caused by total pressure gradients. For mixtures of more than two components,

the Fickian mass transport coefficient becomes an adjustable parameter that must be

obtained by fitting experimental data, even for pore structures that can be represented as

parallel pore bundles. The occurrence of viscous flow, Knudsen flow, Knudsen diffusion,

and bulk diffusion during transport of gases in porous materials is accounted for in the

dusty-gas model (Jackson, 1977; Mason and Malinauskas, 1983; Sotirchos, 1989).

Pore Diffusion Model: The bulk separation of gas mixture was first addressed by

Yang and Doong in 1985 with a pore diffusion model. They studied a 50/50 gas mixture

of methane and hydrogen in activated carbon and, then, extended it for a ternary mixture

of hydrogen, methane, and carbon dioxide (Doong and Yang, 1986). The new model, a

pore diffusion model for mass transfer, was compared to the LDF model by Farooq and

Ruthven (1990). They concluded that the pore diffusion model was complex and

computationally cumbersome and was no better than the simple LDF model.

Thermal Effects: Non-isothermal effects are intrinsic to every sorption process

because of the heat associated with adsorption and desorption. When the process takes

place in small-diameter beds with thick metallic walls, the heat is quickly transported to

the walls where it is stored, and the operation is nearly isothermal despite the presence of

heat effects. In large-diameter beds, such as the ones used commercially, the heat

produced or consumed is not conducted fast to or from the walls, and the temperature

fluctuation in the bed can sometimes be as high as 100 K (Yang, 1987). A comprehensive

discussion and experimental evidence of heat effects in large adsorbent beds is provided

19
by Leavitt (1962). The temperature variation during adsorption and desorption can

dramatically change the performance of the adsorption-based process because the

properties of most of the adsorbents exhibit a strong temperature dependence. Therefore,

the mathematical models used for the design and optimization of adsorption-based

processes must account for the temperature changes in the adsorbent bed. Meyer and

Weber (1967) and Nagel et al. (1987) developed non-isothermal adsorption models in

which both energy equations (for the adsorbent bed and for the adsorbent particle) were

included. However, for adsorption systems with moderate heat effects and moderate

temperature dependence of the adsorption isotherms, the temperature in the particle may

be assumed to be uniform and the energy equation in the particle does not have to be

included in the model (Serbezov and Sotirchos, 1998a). Such non-isothermal models

were proposed by Chihara and Suzuki (1983), Yang and Doong (1985), and Farooq et al.

(1988).

2.5 Multicomponent separation

Separation of more than one impurity from a gas mixture requires more than one

selective adsorbent and, hence, the bed models for multicomponent separation need to

consider different adsorbents. These also create complexity for the inlet conditions for the

second adsorbent bed. Such a model has been studied by several investigators. The

arrangements of adsorbents led two types of adsorption systems: multiple adsorption

column and layered bed adsorption column systems.

Multiple Adsorption Columns: Sircar (1979) patented the first hydrogen

purification system with impurities such as CO2, CO, and CH4. Two adsorption beds

20
were operated in series. A similar approach, a series of beds, was also patented by Kumar

(1990) for the production of hydrogen from coke oven gas by using activated carbon and

zeolite5A. Both of them kept options for independent operation of the beds, which made

the process complicated in terms of operations.

Layered Bed Adsorption Column: Chlendi and Tondeur (1995) were the first to

study fixed-bed adsorption with two layers (activated carbon and molecular sieve 5A) of

adsorbents in a single column for separation of carbon dioxide using an equilibrium

model. Chlendi et al. (1995) extended the previous work for hydrogen purification from

cracked natural gas. They neglected thermal effects for the system and studied the effects

of some operating and design variables on the performance of PSA cycles. Yang and Lee

(1998) studied adsorption dynamics of a layered bed adsorption system with activated

carbon and zeolite5A for hydrogen recovery from coke oven gas. They used a single

composition of the gas mixture and a simplified form of numerical simulation in their

study. They found an intermediate breakthrough behavior.

Later, Lee et al. (1999) extended the study and investigated the effect of the ratio

of bed heights. They found the ratio to affect the separation purity for a given throughput.

Malek and Farooq (1998) studied the removal of hydrocarbons from refinery fuel gas

with a double layered (silica gel and activated carbon) adsorption system. They pointed

out that the heat effect had a significant effect on the performance of PSA cycles.

Methane, ethane, propane, and butane were considered to be major impurities. Jee at el.

(2001) studied the effect of adsorption pressure, feed flow rate, and the ratio of

adsorbents (activated carbon to zeolite 5A) for hydrogen PSA cycles with two adsorbents

21
(activated carbon and zeolite 5A). The notable difference with previous studies was the

inclusion of energy balance.

Takamura et al. (2001) studied a dual bed adsorption system for CO2 separation

from boiler exhaust gas. They also studied the effects of the adsorbent ratio and

concluded that the ratio affects the separation efficiency of the process. Cavenati et al.

(2006) studied separation of CO2 and N2 from a mixture of CH4, CO2, and N2. They used

a layered bed of zeolite13x and CMS 3K in their study. They investigated the

breakthrough dynamics and temperature variation in the bed. They limited their study to

two concentration combinations and a single feed pressure. They also used a fixed

volumetric flow rate in their study. Rebeiro et al. (2008) studied five component

separations in a dual bed of activated carbon and zeolite for hydrogen purification. They

compared a reduced model based on controlling resistance with complete model and

concluded that the effect of micropore resistance was not significant.

Commercial Platforms: The next level of publications on adsorption modeling

included all the features and criticality of adsorption processes together to produce a

general platform with which any process can be explored. Notable examples of such an

approach are Kumar et al. (1994) and Da Silva (1998). Based on these publications,

commercial simulators, such as Aspen Adsorption and gPROM, were built. They offer

various modeling flexibility, which can be customized for a process of interest. Even

procedures for numerical solutions can be chosen.

2.6 Numerical solution of partial differential equations

22
Several numerical methods that address the solution of partial differential

equations (PDEs) with steep fronts and highly non-linear behaviour are available in

literature. For example, Nilchan (1997) and Nilchan and Pantelides (1998) used finite

difference and collocation methods in both time and space to discretize the PDEs, which

were then solved using a non-linear solver in the gPROMS platform. The drawbacks of

the technique include ineffective initialization of a large set of equations and a lack of

guarantee of a real-valued solution (Biegler et al., 2005). Ko et al. (2003) also stated that

complete discretization may lead to convergence failure due to the accumulation of

errors, especially for complicated models. On the other hand, the method of lines (MOL)

is a two-step technique that discretizes the space derivative first and then applies

numerical integration to find approximate solution. The decoupling of the space and time

variable can produce high-order accuracy (Biegler et al., 2005). Ko et al. (2003) found

the technique to be easier and more reliable than complete discretization models.

Discretization of Space Derivatives: Several discretization techniques (finite

difference, finite element, and finite volume) have been applied with first order or higher

order accuracy by different authors (discussed by Beigler et al., 2005) within the MOL

framework. The problem is numerical smearing or oscillation, which were addressed by

introducing a high resolution scheme (Finlayson, 1992), multiresolution scheme (Cruz et

al., 2003) and flux corrected transport (Book, 1981). The flux corrected transport method

has been modified in modern versions. For example, Van Leer used an anti-diffusion step

to avoid excessive smearing. Hirsch (1988), Webley and He (2000), and Jiang et al.

(2003) successfully used several flux limiter methods. The upwind finite differencing

method uses an adaptive or solution-sensitive finite difference stencil to determine the

23
spread of information in a flow field. Historically, the origin of upwind methods can be

traced back to the work of Courant, Isaacson, and Rees (1952) who proposed the CIR

method. It is the preferred option because it is a good all-round performer,

unconditionally non-oscillatory, the cheapest user of simulation time, and reasonably

accurate.

Numerical Integration: The backward differentiation formula (BDF) is a family

of implicit methods for the numerical integration of ordinary differential equations. They

are linear multistep methods that, for a given function and time, approximate the

derivative of that function using information from already computed times, thereby

increasing the accuracy of the approximation. These methods are especially used for the

solution of stiff differential equations. The Gear formulae (Gear, 1971) have great

importance within the multi-step integration methods used in transient processes, since it

allows variable order and variable step change to produce high accuracy.

24
3 Modeling and simulations of a gas adsorber

3.1 Adsorption modeling

A process simulator, namely Aspen Adsorption of Aspen Plus, was used to

simulate the adsorption of CO2 and N2 from natural gas. An adsorption model based on

the following assumptions was formulated in this simulator. The mathematical model

equations together with the correlation used for the estimation of mass and heat transfer

parameters are listed in Table 3.1.

Assumptions for physical adsorption process:

 The flow pattern is described by an axially dispersed plug flow model.

 The mass transfer is described by a lumped overall resistances model.

 The mass transfer driving force is linear and is based on solid film.

 The process is non-isothermal.

 Conductivities (gas, solid, and axial conduction of wall) are negligible.

 Enthalpy of adsorbed phase and enthalpy of mixing are negligible.

 Heat of adsorption is constant.

25
Table 3.1: Model equations

3 2 5
P 1.50  10 (1   i ) 1.75  10 (1   i ) M g 2
 v g 
Ergun Equation
 
vg
2 3 3
z 2 rp  i 2 rpi

Component mass
 i Da
2
 Ci   g Ci


 i
Ci q
 s i  0

balance z
2
z t t

Gas phase energy


Cvg g  g
Tg
  bCvg  g
Tg
p
 g

 h f a p Tg  Ts   
4 H w Tg  T0 0
balance z t z DB

Solid phase energy


 s C ps
Ts n
  s  C paiwi
i1
Ts
 n
  s   H i
i1

qi 
  hf ap Tg  T0   0
balance t t t 

 wC pw
Tw
 Hw
4 DB Tg  Tw
 H amb
4 DB  WT2
Tw  Tamb
0
   
Wall energy balance
t DB  WT
2 2
 DB  DB  WT
2

 DB
2
 
Linear driving force
model
qi
t
*
 k qi  qi 
Mass transfer 1 RpK
2
RpK 2
Rc q   
   where K   0   . s 
coefficient k 3k f 15 p D p 15Dc  C0    i 

 Dm  
0.6 1/ 3 
  g gd p    
Film mass transfer
   2  1.1   
kf
 d p        D  
coefficient
      g m 

 0.6 1/ 3 
Film heat transfer  kg     g  g d p   C pg  

hf
 d    2  1.1    k M  
coefficient  a     g  
 
1
Wall heat transfer hw
 kg 
 
 1.15d p 

 2  10
6
PeH
2
 
 CsphereH B
 0.0477PeH  22.11   1 
  


coefficient    DB PeH 

Axial dispersion Da  0.73Dm   g rp 9.49 i Dm / 2 g rp   1


1/ 2
3 1.75 

1 1 

1.00  10 T
M  M 
Diffusion equation
D AB 
 A B
Dk  9700rmac
T

 
,
1/ 3 2
P  A Vi
1/ 3
  B Vi    M

 
  
1 1 1
Bosanquet equation 
D pi  D eff D eff

 m,i  k ,i

Note: References for equations are included in Section 3.1.1 Model equations.

26
3.1.1 Model equations

Momentum balance equation: The Ergun equation, which combines the

description of pressure drops in the Karman-Kozeny equation for laminar flow and the

Burke-Plummer equation for turbulent flow, is useful for both laminar and turbulent flow

(Bird et al., 1960).

P 1.50  10 3 ( 1   i )2 1.75  10 5 ( 1   i )M g 2


  
2rp 2  i 3
v vg (3.1)
z 2rp i
g 3

where P is feed pressure (bar), z is the height of the adsorbent bed, εi is the interparticle

Voidage, νg is the gas velocity (m/s), ψ is the shape factor, μ is the dynamic viscosity (N

s/m2), M is the molecular weight (kg/kmol), and rp is the particle radius (m).

Mass balance equation (ref: Aspen manual): The transfer of mass from gas to a

solid surface can occur in four ways: dispersion, convection, accumulation, and diffusion.

Dispersion may happen in radial or axial directions. Radial dispersion was not considered

in this model.

 2Ci  g Ci  Ci q


 i Da   b  s i  0 (3.2)
z 2
z t t

where Ci is the molar concentration of component i (kmol/m3), Da is the axial dispersion

coefficient of component i (m2/s), εi is the interparticle voidage, εb is the bed voidage, t is

the time (second), ρs is the adsorbent bulk density (kg/m3), and qi is the loading

(kmol/kg).

Gas phase energy balance (ref: Aspen manual): The gas phase energy balance

includes the convection of energy, accumulation of heat, compression, heat transfer from

gas to solid, and heat transfer from gas to the internal wall. Conductive heat transfer of

gas has not been considered.

27
Tg Tg  g
 h f a p Tg  Ts   Tg  T0   0
4H w
Cvg g  g   B Cvg  g p (3.3)
z t z DB

where Cvg is the specific gas phase heat capacity at constant volume (MJ/kmol/K), ρg is

the gas phase molar density (kmol/m3), Tg is the gas phase temperature (K), Ts is the solid

phase temperature (K), To is the ambient or wall temperature according to context use

(K), hf is the gas-solid heat transfer coefficient (MW/m2/K), ap is the specific particle

surface per unit volume bed (m2 (Particle area)/m3 (Bed)), Hw is the gas-wall heat transfer

coefficient (MW/m2/K) and Db is the bed diameter (m).

Solid phase energy balance (ref: Aspen manual): The solid phase energy balance

includes the accumulation of heat, accumulation of enthalpy in the adsorbed phase, heat

of adsorption, and gas-solid heat transfer from gas to solid through the film at the solid

surface. The heat transfer area was assumed to be proportional to the area of the

adsorbent particles. The conductive heat transfer from the solid was not considered for

this application.

Ts T q 
  s  C paiwi  s   s   H i i   h f a p Tg  T0   0
n n

 s C ps (3.4)
t i 1 t i 1  t 

where Cps is the specific heat capacity of adsorbent (MJ/kmol/K), Cpai is the specific heat

capacity of the adsorbed phase (MJ/kmol/K), wi is the solid phase concentration

(kmol/kg), and ΔHi is the heat of adsorption of component i (MJ/kmol).

Wall energy balance (ref: Aspen manual): The wall energy balance includes heat

accumulation within the wall material, heat transfer from the bed to the inner wall, and

heat transfer from the outer wall to the environment.

Tw 4 DB Tg  Tw  4DB  WT  Tw  Tamb 


2
 wC pw  Hw  H 0 (3.5)
t DB  WT 2  DB2 amb
DB  WT 2  DB2

28
where, Cpw is the specific heat capacity of column wall (MJ/kg/K), ρw is the wall density

(kg/m3), WT is the width of column (m), Tw is the wall temperature (K), Tamb is the

ambient temperature (K), and Hamb is the wall-ambient heat transfer coefficient

(MW/m2/K).

Linear driving force (LDF) model: The concentration difference between bulk

gas and the solid phase sets the mass transfer driving force (mass transfer coefficient).

One model that has found wide and successful use in the analysis and design of

adsorptive separation processes is known as the linear driving force (LDF) model

(Alvarez-Ramirez et al., 2005). This first order model was proposed by Gleuckauf and

Coates (1947) for adsorption chromatography. The model is simple, analytical, and

physically consistent (Sircar and Hufton, 2000) and realistically represents an industrial

process (Biegler et al., 2005).

qi
t

 k qi*  qi  (3.6)

where q is the average adsorbate concentration in the solid (kmol/kg), q* is the

instantaneous equilibrium concentration, and k is the effective mass transfer coefficient of

component i.

Isotherm Model (ref: Aspen manual): The temperature-dependent Langmuir-

Freundlich isotherm (adsorption equilibrium at fixed temperature) model was used in this

study. This isotherm takes the advantages of monolayer adsorption, described by the

Langmuir model for low pressure, and multilayer adsorption, described by the Freundlich

model for high-pressure adsorption. The multicomponent adsorption equilibria were

introduced through the ideal adsorption solution theory (IAS). The Langmuir-Freundlich

29
model for a temperature-dependent multicomponent system was presented in Aspen

Adsorption as:

1
 a   a 
 4   6 
 a T   a T 
qi   a  a  P 3  e s   1  a  P 3  e s  (3.7)
 1 2 i   5 i 
   
   

where a1, a2, a3, a4, a5, and a6 are constants and Pi is the partial pressure of component i.

Mass transfer coefficient: There are three mass transfer resistances in an

adsorption column: film resistance, macropore resistance, and micropore resistance.

Through a moment analysis of the pulse response from a chromatographic column model,

Haynes and Sharma (1975) came up with following equation. The mass transfer

coefficient obtained from this equation was used in the LDF model.

1 Rp K R p2 K R2
   c (3.8)
k 3k f 15 p D p 15Dc

where kf is the film mass transfer coefficient (m/s), Dp is the pore diffusivity/macropore

diffusivity (m2/s), Rc is crystal radius (m), Dc is crystal diffusivity/micropore diffusivity

(m2/s), and K is dimensionless Henry’s Constant.

For the nonlinear system, it works with reasonable accuracy when Henry’s

constant is replaced by q0/C0, where C0 is the feed concentration of the adsorbate in the

gas phase, and q0 is the corresponding equilibrium in the adsorbed phase (Hassan et al.,

2008). The constant (K) must be in a dimensionless form:

 s  q0    s 
K  KH     . 
 i  C0    i  (3.9)

30
where KH is Henry’s constant, q0 is initial solid phase concentration, and C0 is initial bulk

concentration.

Film mass transfer coefficient: The mass transfer coefficient for a stagnant film

surrounding a solid adsorbent packed inside a fixed-bed adsorber was calculated from the

following correlation (Wakao and Funazkri, 1978):

Sh  2  1.1 Re0.6 Sc1 / 3 (3.10)

k d 
where Sh is the Sherwood Number  f p  , dp is the particle diameter, Dm is the
 Dm 

molecular diffusivity, kf is the film mass transfer coefficient, Sc is the Schmidt Number

   vd 
  and Re is the Reynolds Number  g p g 
 .
 D   
 g m  

The calculation of this film mass transfer coefficient requires physical properties

of fluids. For a gas mixture, the following correlations (Griskey, 2002) were used to

determine mixture properties. The correlation of Fuller et al. (1966) was used for the

calculation of binary diffusivity, which works for both polar and non-polar molecules.

The diffusivities were then corrected for tortuosity as suggested by Do (1998) and then

component diffusivity was calculated using Wilkes’ formula (Equation – 3.13; as

explained by Do, 1998)

n
 mix    i yi (3.11)
i 1

n
yi  i
 mix   n
i 1
y
j 1
j ij
(3.12)

31
2
1  M i 
1 / 2
   1/ 2  M 1/ 4 
ij  1 1   i   j   (3.13)
2 2  M j     j   M i  
 

1/ 2
 1 1 
1.00  10 T 3 1.75
  
DAB   MA MB  (3.14)
P  V    V  
A i
1/ 3
B i
1/ 3 2

p
Corr
DAB  D (3.15)
 2 A ,B
1
 n 
 yB 
Dmeff,i  (1  y A )  corr  (3.16)
 B1 DA,B 
 B A 

where ρmix is the density of the gas mixture, μmix is the viscosity of the gas mixture, yi is

the mole fraction of component i and φij the viscosity of component i and j, DAB is the

binary diffusivity (cm2/sec) of component A and B, Vi is the atomic diffusion volumes

(cm3), Dmeff,i is the effective component molecular diffusivity, DAcorr


, B is corrected binary

diffusivity, and τ is tortuosity

Pore diffusion in macropore: Diffusion in gases is the result of the collision

process (Bird et al., 1960). In a macropore, two types of collision take place, which

results in two types of diffusivities: molecular diffusivity (collision between molecules)

and Knudsen diffusivity (collision between molecules and the pore wall). Depending on

the mean free path of gas molecules and pore diameter, one diffusion process may

dominate over others. The effect can be qualitatively predicted by using the concepts of

the Knudsen number and mean free path as outlined by Do (1998). Molecular diffusivity

has already been discussed in a previous section. Knudsen diffusivity can be calculated

using Equation 3.17 as presented by Smith (1970) and is valid for a capillary tube.

32
Equation 3.18 was suggested by Do (1998) for effective Knudsen diffusivity in porous

media. Pore diffusivity (Equation 3.19) is the combined effect of molecular and Knudsen

diffusivity. It can be calculated using the Bosanquet Equation (Cavenati et al., 2006):

T
Dk  9700 rmac (3.17)
M

b
Dkeff  D (3.18)
2 k

1  1 1 
   eff  eff 
 (3.19)
D pi  Dm ,i Dk ,i 

where Dk is the Knudsen diffusivity, rmac is the macropore radius, Dkeff is the effective

Knudsen diffusivity, and Dpi is pore diffusivity of component i.

Axial dispersion: The axial dispersion coefficient (Da) varies along the length of

the bed. Aspen Adsorption estimates the values during the simulation using the following

correlation:

Da  0.73Dm   g rp 9.49 i Dm / 2 g rp 


1
(3.20)

Film heat transfer coefficient: The correlation published by Wakao and Funazkri

(1978) was used in the calculation of the film heat transfer coefficient. This equation also

counts for dispersion effects:

Nu  2.0  1.1Re0.6 Pr1/3 (3.21)

h d   C pg 
where Nu is the Nusselt number  f a  , Pr is the Prandtl number   and da is the
 k  k M 
 g   g 

diameter as an agglomerate (Do, 1998).

Gas-Wall heat transfer coefficient: The determination of gas-wall heat transfer is

critical since the exact nature of contact between solid particles and the wall is unknown.

33
Kast (1988) graphically represented the relationship. The following correlation uses

results from the graphical representation given by Kast (1988). The value obtained has

been used as model input.

1
 kg 
hw  
 1.15d 
   C

H 
   2  10 6 Pe H 2  0.0477 Pe H  22.11  1  sphere B 
DB Pe H 
(3.22)
 p  

where Csphere is 12 for a packed bed of spheres, PeH (1.15dpνgρgMCpg/kg) is the Péclet

number for gas wall heat transfer, and kg is the conductivity of the gas phase (MW/m/K).

Wall-ambient heat transfer coefficient: This coefficient should follow the exact

environment in which the experiments were performed. Air or water is generally used as

an external cooling medium and the heat transfer coefficient for air and water are 10-100

W/m2/K and 500-10000 W/m2/K, respectively.

Heat of adsorption: The heat arising due to the adsorption of a certain amount of

molecules is known as the heat of adsorption. Though many forms of heat of adsorption

have already been discussed in the literature, isosteric heat of adsorption directly

describes the non-isothermal behavior of adsorption systems (Sircar et al., 1999). This

key thermodynamic variable plays an important role in the design of adsorption systems

because it changes the adsorbent temperature.

3.1.2 Solution of model equations

The model equations were solved numerically using the method of line (MOL)

technique, a built-in method in Aspen Adsorption. This two-step technique discretized

the space derivative first and then applied numerical integration to find an approximate

solution. The space derivatives were set to discretize by using the upwind finite

34
difference method. The method offers good all-round performance. Numerical

integrations were performed using Gear formulae (Gear, 1971). The formulae use the

implicit backward differentiation technique.

3.1.3 Calculation procedure

The basic inputs required for starting the calculations are: the physical properties

of the column and adsorbents, feed conditions, mass transfer, heat transfer, and isotherm

parameters. Some of these inputs were fixed while some others varied with operating

conditions. For example, the porosity of adsorbent was fixed during the study, while

mass transfer parameters changed with operating conditions. The mass and heat transfer

parameter can be obtained using equations described in section 3.1.2. These equations

also require the physical properties of feed gases. In this study, the NIST (National

Institute of Standard and Technology) database was used for obtaining the physical

properties of CH4, CO2, and N2. Aspen Adsorption’s NIST thermo engine has access to

this database.

Figure 3.1 shows the inputs for a single-bed adsorption column. A layer bed

column, for example, with 2 layers of different adsorbent requires 2 sets of properties for

adsorbents. The introduction of a two-adsorbents layer will also influence the overall

flow passage for the feed. The number of parameters for the isotherm depends on the

isotherm model. For example, the Langmuir model is comprised of two parameters, while

the Langmuir-Freundlich model is comprised of six parameters for each component of

feed. Aspen starts the calculation assuming an initial velocity. A dynamic run produces

product composition in every time interval. The feed and product information was then

35
processed to determine the amount of adsorbent and separation efficiency, which are

defined below.

 Amount of adsorbent refers to the following relationship:

M  kg 
Amount of adsorbent, w    (3.23)
F t  mol 

where, M is the mass of adsorbent (kg), F is the feed flow rate, and t is the cycle time (s).

 Separation efficiency refers to the amount of nitrogen or carbon dioxide adsorbed

in the adsorbent. It was defined as:

nin  nout
Separation efficiency,  s   100 (%) (3.24)
nin

where n is the number of moles of nitrogen or carbon dioxide.

36
Input Simulator Output

Feed
Pressure
Temperature
Composition
Flow rate
Column
Radius
Porosity
Specific heat Product
Wall thickness Flow rate
Aspen Adsorption
Adsorbent Pressure
Bed length Temperature
Bulk density Composition
Bulk porosity
Extrude radius
Extrude density Post Processing
Extrude porosity Amount of adsorbents
Transport Parameters Separation efficiency
Mass Transfer coefficient Correlations
Heat transfer coefficient
Heat of adsorption
Specific heat
Thermal conductivity
Isotherm parameters

Figure 3.1: Calculation procedure

37
3.2 Model validation

The model was validated against three different separation systems that covered

pure and multicomponent feeds, homogeneous and heterogeneous adsorbents, isothermal

and non-isothermal heat balance, a wide pressure range, and a wide range of

concentrations of adsorbate to ensure the versatility of the model. The systems were as

follows: N2 separation from helium using activated carbon as per Shen et al. (2010), CH4

separation from hydrogen using Zeolite 5A as per Bastos-Neto et al. (2010), and CO2

separation from a mixture of CH4, CO2, and N2 using zeolite13X as per Cavenati et al.

(2006). The operating boundaries, such as feed pressure, for these three systems varied

widely (1.0-20.2 bar), though the temperature range is quite narrow (300-303 K). The

concentration of adsorbate in feed varied from 0.5 to 20 mol %. This operating window

reflected the likely feed conditions such as pressure, temperature, and composition of a

typical natural gas transmission pipeline.

3.2.1 Nitrogen separation using activated carbon

Shen et al. (2010) reported several experimental breakthrough behaviours of N2 in

pitch-based activated carbon beads (ACB) in diluted conditions (0.5% N2 and 0.995%

helium) at 1 bar and for a wide range (303-423 K) of temperatures. The breakthrough

curve at 303K, as it is close to the temperature of natural gas pipelines, was selected for

comparison and was compared with the simulated breakthrough curve obtained from the

proposed model. The inputs for the model are summarized in Table 3.2. The physical

properties of the adsorbent were taken from the literature. The mass transfer coefficient

was calculated and isotherm parameters were obtained from the fitting of the Langmuir

38
isotherm, which fitted the experimental data with an average absolute deviation (AAD) of

1.94%. Only two constants of the isotherm model were used because the separation

process represents single component separation in a diluted condition. Heat transfer

parameters were omitted from the input table as the process was isothermal. Figure 3.2 is

a comparison of the experimental and simulated results of authors to model the output of

this work. The plots portray the accuracy of the model in terms of both, shape and width

of the breakthrough curves. The AAD between experimental data and model output (our

study) is 3.1% and that of their model is 2.9%. This thereby validates the model.

39
Table 3.2: Model input (N2-ACB system)

Descriptions Value
Pressure (bar) 1
Temperature (K) 303
Concentration of nitrogen (% mole) 0.5
Concentration of helium (% mole) 95.5
Height of adsorbent layer (m) 0.165
Internal diameter of adsorbent layer (m) 0.0093
Inter-particle voidage (m3 void/m3 bed) 0.32
Intra-particle voidage (m3 void/m3 bed) 0.51
Bulk solid density of adsorbent (kg/m3) 669
Adsorbent particle radius (m) 5.45E-04
Adsorbent shape factor ( ) 1
Constant mass transfer coefficients (1/s) 0.48
Isotherm parameter, a1 (n/a) 3.31E-04
Isotherm parameter, a2 (n/a) 0.0977
Note: Langmuir isotherm model was used.

40
1.00
Mole ratio of N2 at bed exit (C/C0 )

0.80

0.60

Experiment
0.40
(Shen et al.,
2010)
Simulation
0.20
(This Study)

0.00
0 100 200 300 400
Time (second)

Figure 3.2: Breakthrough concentration profiles of N2 in pitch-based AC beads (0.5% N2

in helium at 303K and 1 bar) under isothermal conditions

41
3.2.2 Methane separation from hydrogen using zeolite 5A

Bastos-Neto et al. (2010) presented a series of adsorption isotherms of CH4 and

breakthrough curves on one zeolite 5A at various concentrations, temperatures, and

pressure conditions. They used a mixture of CH4 and H2 to analyze the recovery of CH4

from H2. From those breakthrough curves, a high pressure at 20.2 bars on the

breakthrough curve was selected to check the high-pressure response of our model and to

determine the CH4 handling capability of the model at a moderate concentration of

adsorbate (8.8 mol % CH4, balance H2) and high pressure. Inclusion of zeolite 5A also

helped to test various adsorbent handling capability of the model.

The inputs of the model are summarized in Table 3.3. Heat transfer parameters

are not included in Table 3.3 because the system is isothermal. Physical properties of

adsorbents and column dimensions were obtained from the literature to match the

experimental arrangements used by Bastos-Neto et al. (2010). The experimental

adsorption data obtained from the literature were fitted with the Langmuir-Freundlich

isotherm. The isotherm model fitted the experimental adsorption data with an AAD of

3.23%. Mass transfer parameters were calculated using the correlations mentioned in this

work (Section 3.1.2).

Figure 3.3 portrays experimental and simulated breakthrough profiles. The shape

and width of the breakthrough curve proved the accuracy of our model. The average

absolute deviation between experimental data and model output were 5.75%. This

comparison especially highlights the capability of the model in handling moderate

concentrations of contaminant such as 8.8 mole percent CH4 in the feed at high pressure.

42
Table 3.3: Model inputs (CH4-H2-Zeolite5A system)

Descriptions Value
Pressure (bar) 20.2
Temperature (K) 300
Concentration of methane (% mole) 8.8
Concentration of hydrogen (% mole) 91.2
Height of adsorbent layer (m) 0.08
Internal diameter of adsorbent layer (m) 0.022
Inter-particle voidage (m3 void/m3 bed) 0.44
Intra-particle voidage (m3 void/m3 bed) 0.41
Bulk solid density of adsorbent (kg/m3) 706
Adsorbent particle radius (m) 0.001
Adsorbent shape factor 1
Constant mass transfer coefficients (1/s) 0.01
Isotherm parameter, a1 (n/a) 7.8E-04
Isotherm parameter, a2 (n/a) 0.1895
Note: Data were obtained from Bastos-Neto et al. (2010).

43
1.00
Mole ratio of CH4 at bed exit (C/C0 )

0.80

0.60

Simulation (This
0.40 Study)

Experiment
0.20 (Bastos-Neto et al.
(2010))

0.00
0 1200 2400 3600 4800 6000
Time (second)

Figure 3.3: Breakthrough concentration profile of methane in Zeolite 5A (8.8% Methane


in Hydrogen at 303 K and 20.2 bar) under isothermal conditions

44
3.2.3 Carbon dioxide separation using Zeolite 13X

The separation of CO2 from a mixture of CH4, CO2, and N2 was studied by

Cavenati et al. (2006) using zeolite13X. This ternary gas mixture (70 mol % CH4, 20 mol

% CO2, and 10 mol % N2) reflects the likelihood of components as well as their

concentrations in natural gas and enforces the inclusion of multicomponent non-

isothermal behavior, a more practical phenomenon in gas separation industries, in the

model. The study was performed at three different temperatures (298K, 308K, and

323K). A temperature-dependent isotherm model, the Langmuir-Freundlich, was used to

relate the equilibrium data. The breakthrough behaviors of the system were evaluated

using the data provided by the authors.

The inputs used for this system are summarized in Table 3.4. Isotherm parameters

are in Table 3.5. Some transport properties were calculated using the

equations/correlations discussed in section 3.1.1. The results of our model were presented

graphically in Figures 3.4 – 3.5. The column dynamics, i.e., the breakthrough and

temperature profiles, were compared in Figures 3.5 – 3.6. Breakthrough dynamics differ

by 2.1% (AAD), while temperature profiles differ by 0.7% (AAD) when compared to the

experimental data. The cooling side of the temperature profile notably differs from the

experimental data. This might be due to the end effect on the temperature probe that was

placed between two layers of adsorbent. The model fairly describes the non-isothermal

behaviour associated with the removal of CO2 with a relatively high concentration. It is

also capable of handling multicomponent gas and the temperature dependency of the

isotherm.

45
Table 3.4: Model inputs (CH4-CO2-N2-zeolite13X system)

Description Value
Pressure 2.5
Temperature 300
Height of adsorbent layer (m) 0.20
Wall thickness (m) 0.0024
Internal diameter of bed (m) 0.016
Inter-particle voidage 0.33
Intra-particle voidage 0.46
Bulk density (kg/m3 ) 715.00
Particle radius (m) 9.00E-04
Specific heat capacity (J/kg/K) 880.00
Wall specific heat capacity (J/kg/K) 500.00
2
Heat transfer coefficient (W/m /K)
overall 67.00
gas to wall 192.00
wall to ambient 37.00
Wall thermal conductivity (W/m/K) 13.40
3
Wall density (kg/m ) 8238.00
Mass transfer coefficients (1/s)
CH4 3.49E-05
CO2 7.26E-03
N2 4.74E-03
Heat of adsorption (MJ/Kmol)
CH4 -38.95
CO2 -38.95
N2 -15.93

46
Table 3.5: Isotherm parameters (CH4-CO2-N2-zeolite13X system)

Component Parameter Value


a1 1.71E-04
a2 1.72E-04
a3 7.16E-01
CH4
a4 3.19E+03
a5 1.06E-05
a6 2.98E+03
a1 4.98E-04
a2 1.05E-02
a3 4.92E-01
CO2
a4 2.16E+03
a5 1.47E-03
a6 2.04E+03
a1 3.37E-06
a2 8.71E-01
a3 7.96E-01
N2
a4 1.46E+03
a5 1.06E-03
a6 1.36E+03

47
1.00
Mole Ratio of CH 4, CO 2 and N2 at bed exit (C/C0)

0.75

CH4 (Experiment by Cavenati et al., 2006)


Simulation,CH4 (This Study)
0.50 CO2 (Experiment by Cavenati et al., 2006)
Simulation, CO2 (This study)
N2 (Experiment by Cavenati et al., 2006)
Simulated, N2 (This Study)
0.25

0.00
0 400 800 1200 1600 2000 2400
Time (seconds)

Figure 3.4: Breakthrough concentration profiles of 70 mol % CH4, 20 mol % CO2 and 10
mol % N2 in Zeolite 13X at 300K and 2.5 bars

48
335

330
Temperature at bed exit (K)

Simulation (This
325 study)
Experiment (Cavenati
320 et al., 2006)
Simulation (Cavenati
315 et al., 2006)

310

305

300

295
0 400 800 1200 1600 2000 2400
Time (second)

Figure 3.5: Temperature profile at bed exit (70 mol % CH4, 20 mol % CO2 and 10 mol %
N2 in Zeolite 13X at 300 K and 2.5 bars)

49
4 Results and Discussion

4.1 Description of simulated gas adsorption systems

This work simulated the adsorption of CO2 and N2 from natural gas containing

50-90% CH4, 5-25% CO2, and 5-25% N2 in layered bed adsorbers. As depicted in Figure

4.1, the layered bed adsorber contains two layers of selected adsorbents, namely

zeolite13X and activated carbon (ACB) or a carbon molecular sieve (CMS3K). All these

adsorbents can adsorb CH4, CO2, and N2. However, zeolite13X has greater adsorption

capacity with CO2 than CH4 and N2 (Cavenati et al., 2004). ACB and CMS3K also

preferentially adsorb CO2 over N2. Thus, zeolite13X is placed at the bottom of the

adsorber where the gas feed was introduced to completely remove CO2 and ACB, or

CMS3K is placed at the top to remove N2.

A single column was filled with two adsorbents. First, the amount of zeolite13X

was adjusted to have a CO2 breakthrough at the desired cycle time. Then, amount of

ACB/CMS3K were adjusted in steps to obtain a nitrogen separation efficiency of over

90%. Two cycle times were used: 80 seconds for zeolite13X-CMS3K to implement

kinetic preference of CMS3K for N2 over CH4 and 600 seconds for zeolite13X-ACB to

implement equilibrium adsorption of N2 in ACB. The transport parameters were adjusted

(Appendix – A) to maintain a simplified model without sacrificing the rigour of the

model. Adsorption equilibrium information and physical properties of zeolite13X and

CMS3K were obtained from Cavenati et al. (2004 and 2006) and that of ACB were

obtained from the work of Shen et al. (2010).The heat transfer coefficient between the

wall and environment was obtained from the work of Cavenati et al. (2006).

50
Product

Feed

Figure 4.1: Double bed adsorber

51
4.2 Simulation results for zeolite13X

Simulations were carried out on single-bed zeolite13X for different feed pressures

and compositions for complete separation of CO2. Five different feed pressures (2.5 bars,

5 bars, 10 bar, 20 bars, and 30 bars) were analyzed. Concentration of CO2 and N2 were

varied from 5 to 25%. The amount of CH4 was kept constant at 70%. The inputs are

presented in Tables 4.1 and 4.2. These inputs were for feed containing 70% CH4, 20%

CO2, and 10% N2 at 2.5 bars and 300K. The feed flow rate was 1.6 SLPM. Some of the

inputs, for example mass transfer coefficients, were calculated for every concentration

and every pressure. The isotherm parameters were obtained by fitting the experimental

adsorption data (Cavenati et al., 2004) and were kept unchanged. Physical properties of

zeolite13X were obtained from the work of Cavenati et al. (2006). The wall to ambient

heat transfer coefficient was adjusted to obtain the temperature profile published by

Cavenati et al. (2006).

The height of the adsorber bed was varied to get the breakthrough of CO2 at the

desired cycle time. The breakthrough implies the optimum amount of zeolite13X needed

for 100% separation of CO2. The amount (gram per mole of feed) is presented in Figure

4.2 as a function of feed pressure and concentration of CO2.

52
Table 4.1: Physical properties of zeolite13X and properties of column

(Cavenati et al., 2006)

Column Zeolite13X
Height of adsorbent layer/column, m 0.2 0.2
Wall thickness of column, m 0.0024
Internal diameter of column, m 0.016
Density of column wall, kg/m3 8238
Inter-particle voidage/column porosity 0.33 0.33

Specific heat capacity, J/kg/K 500 920


Bulk density/column density, kg/m3 756 756
Thermal conductivity of column wall, W/m/K 13.40
Particle/extrude density of zeolite13X 1130
Particle/extrude radius of zeolite13X, m 0.008
Intra-particle voidage/extrude porosity 0.54
Extrude tortuosity 2.2

53
Table 4.2: Parameters used in simulation for zeolite13X

Total Pressure, bar 2.5, 5, 10, 20, 30


Flow rate, SLPM 1.6
Temperature, K 300
Mass transfer coefficient, 1/s CH4 – 0.3315
CO2 – 0.0165
N2 – 0.5053
Heat transfer coefficient, W/m2/K Gas-Solid – 60
Gas-wall – 36
Wall-ambient – 192
Isotherm parameters for CH4 a1 – 0.0112
a2 – 0.00029
a3 – 0.834
a4 – 1590
a5 – 0.00037
a6 – 1610
Isotherm parameters for CO2 a1 – 0.00033
a2 – 0.0021
a3 – 0.43
a4 – 2850
a5 – 0.0031
a6 – 2450
Isotherm parameters for N2 a1 – 0.0022
a2 – 0.00042
a3 – 0.885
a4 – 1740
a5 – 0.0001
a6 – 1970
Note: Transport coefficients were calculated (70%CH4/ 20%CO2/10% N2/2.5bar/300K).

54
100
90
80
Amount of zeolite13X

70
(g/mol of feed gas)

60
50
40
30 05%CO2 - 25% N2
10%CO2 - 20% N2
20 15%CO2 - 15% N2
20%CO2 - 10% N2
10 25%CO2 - 05% N2
0
0 5 10 15 20 25 30
Feed gas pressure (bar)

Figure 4.2: Required amount of zeolite13X for complete separation of CO2 as a function

of feed pressure

55
4.2.1 Parametric study

Parametric studies were performed to determine the effect of feed conditions on

adsorption of CO2 and N2. Two effects were determined. They are (i) the effect of feed

pressure and (ii) the effect of concentration.

Effect of feed pressure: As per figure 4.2, CO2 adsorption capacity of zeolite13X

increases in low pressure ranges (2.5 to 5 bars), while capacity decreases in high pressure

ranges (10 to 30 bars). This reversal of adsorption capacity occurs for feed pressure

ranging from 5 to 10 bars. The effect of feed pressure becomes less significant at the

pressure range of 20 to 30 bars. The possible reason behind this reversal of adsorption

capacity could be multicomponent adsorption in zeolite13X; more specifically, selective

adsorption of CO2 over N2 might play the significant role. To further investigate this, the

adsorption capacities presented in Figure 4.3b were determined using the Langmuir-

Freundlich model, while the selectivity (Figure 4.3a) was obtained from the graphical

representation. As expected, N2 adsorption capacity of zeolite13X is low (0.05 mol/kg)

for a low partial pressure (0.13 bar) and high (1.44 mol/kg) for a high partial pressure

(7.50 bar) of N2. However, selective adsorption of CO2 over N2 is high (37.1) at a low

partial pressure (0.13 bar) and low (0.30) at a high partial pressure (7.50 bar) of CO2.

This reversal of selectivity implies that N2 was preferentially adsorbed in zeolite13X at a

high concentration of N2. The increased adsorption of N2 on zeolite13X is presented in

Figure 4.5 as a function of feed pressure. As presented, less of 10% (mole) of total N2

was adsorbed at 2.5 bars, while the amount exceeds 40% at 30 bars.

56
40

35

Selectivity of CO 2 over N2
30

25

20

15

10

0
0 1 2 3 4 5 6 7 8
Partial pressure of CO2 (bar)

a) Selectivity of CO2 over N2

7
Adsorption capacity of CO2 or N2 (mol/kg)

4 CO2
N2
3

0
0 1 2 3 4 5 6 7 8

Partial pressure of CO2 or N2

b) CO2 or N2 adsorption capacity (mol/kg) of zeolite13X

Figure 4.3: Adsorption capacities and selectivity for CO2-N2-zeolite13X system

57
Effect of concentration of CO2 and N2: The amount of zeolite13X needed for

100% separation of CO2 is presented in Figure 4.4. Since the concentration of CH4 was

fixed, a low concentration of CO2 implies a high concentration of N2 and vice versa. The

amount of zeolite13X increases linearly for all concentrations of CO2 at 2.5 bars. The

trend is similar for 5 bars but deviates from linearity in 10, 20, and 30 bars. When

compared to the amount of required at 2.5 bars, the required amount of zeolite13X is less

at 5 bars for all concentrations of CO2. At 10 bars, the trend is true for higher

concentrations (above 10%) of CO2. A high amount of zeolite is needed, when compared

to 2.5 to 10 bars, at 20 bars and above for low concentrations (less than approximately

15%) CO2. The behaviour can be explained in terms of selective adsorption, as explained

before (see the effect of feed pressure). Selective adsorption of N2 over CO2 implies that

more N2 will be adsorbed. Figure 4.5 shows that high N2 efficiency was obtained at high

feed pressure and high concentration of N2.

58
100
90
80
Amount of zeolite13X

70
(g/mol of feed gas)

60
50
40
30 Feed pressure
20 2.5 bar 5.0 bar
10.0 bar 20.0 bar
10
30.0 bar
0
0 5 10 15 20 25 30
Concentration of CO2 (%)

Figure 4.4: Effect of concentration (%) of CO2 on required amount of zeolite13X for

100% separation of CO2

59
50
45
40
N2 separation efficiency (%)

35
30
25
20
15 25% N2
20% N2
10 15% N2
10% N2
5 5% N2

0
0 5 10 15 20 25 30 35
Feed pressure (bar)

Figure 4.5: N2 separation efficiency of zeolite13X

60
4.2.2 Correlation to determine amount of zeolite13X

The amounts of zeolite13X determined from the feed pressures were correlated in

a single equation:

C1  (C 2  C 3  P C 4 )  y C 5 (1  y ) C 6
Q 4.1
C 7  C8  P C 4  y C 9  (1  y ) C10

where Q is the amount of adsorbent, P is the feed pressure, y is the mole fraction of N2 or

CO2, and C1 to C10 are parameters. The correlation determines the total amount of

zeolite13X depending on feed pressure. The parameters (C1 to C10) are shown in Table

4.3. There are 2 sets of parameters listed in this table for two pressure ranges. The mole

fraction (y) in the correlation refers to the mole fraction of CO2 in the feed stream, and

the separation factor (s) in the correlation was 1 as the zeolite13X removed all (100%)

carbon dioxide from the feed.

Figure 4.6a presents a comparison of the predicted results for low feed pressures

(2.5 to 10 bar). The average absolute deviation (%) that was observed for this feed stream

was 1.3%. The maximum absolute deviation was 5.6% for the feed stream that carried

05% carbon dioxide at 5 bars. Figure 4.6b compares the predicted results for high feed

pressures (10 to 30 bar). The average absolute deviation (%) that was observed for this

feed stream was 0.36 %. Maximum absolute deviation was observed for the feed stream

that carried 10% carbon dioxide at 10 bars. The predicted amount of zeolite13X in the

feed conditions was lower by 1.15 wt. %.

61
Table 4.3: Parameters of correlation for determination of amounts of zeolite13X

Feed pressure
Parameter
<10 bar ≥ 10 bar
C1 1.559 2.639
C2 404.474 -3.344
C3 0 2.653
C4 13.203 0.84
C5 0.505 -0.107
C6 -0.61 -7.951
C7 3.843 1.188
C8 0 0.111
C9 0.354 0.107
C10 1.01 -7.014

62
100

Predicted amount of zeolite13X


80

(g/mol of feed)
60

40
5% CO2
10% CO2
20 15% CO2
20% CO2
25% CO2
0
0 20 40 60 80 100
Simulated amount of zeolite13X
(g/mol of feed)

(a) Feed pressure 2.5 to 10 bars

100
Predicted amount of zeolite13X

80
(g/mol of feed)

60

5% CO2
40
10% CO2
15% CO2
20 20% CO2
25% CO2

0
0 20 40 60 80 100
Simulated amount of zeolite13X
(g/mol of feed)

(b) Feed pressure 10 to 30 bars

Note: Simulated means results obtained from Aspen Adsorption while


predicted means results obtained from empirical correlation

Figure 4.6: Comparison of simulated and predicted amounts of zeolite13X

63
4.3 Simulation results for zeolite13X-CMS3K system

Simulations were performed for various feed conditions using a layered bed of

zeolite13X-CMS3K. The obtained data were then processed to determine the amount of

adsorbents and adsorbed N2. The system adsorbed all CO2 in zeolite13X. Four pressure

stages (2.5 bars, 5 bars, 7.5 bars, and 10 bars) were studied using 15-25% N2, 10-20%

CO2, and 55-75% CH4. The concentrations of N2 and CO2 were varied in 10% intervals.

The study was performed for a cycle time of 80 seconds.

Sample inputs for the system are presented in Tables 4.4 and 4.5. The amount of

zeolite13X was adjusted to adsorb all the CO2 from the feed before starting the

simulation of the dual bed adsorption column. The amount of this zeolite13X was kept

fixed for particular feed condition, while the amount of CMS3K was varied to get a N2

separation efficiency ranging from 50 to 95%. No attempt was made for 100% removal

of N2. Since all CO2 was adsorbed in zeolite13X, this discussion will be limited to N2

separation efficiency only. The total amount of adsorbents needed for the removal of CO2

and N2 has been plotted in Figure 4.7. Figure 4.7 shows amount of zeolite13X and

CMS3K as function of feed pressure and concentration and N2 separation efficiency. An

increase in N2 separation efficiency required more CMS3K while an increase in feed

pressure required less CMS3K. The effects of concentration of N2 were not significant.

64
Table 4.4: Physical properties of double bed adsorber (zeolite13X-CMS3K)

(Cavenati et al., 2006)

Column Zeolite13X CMS3K


Height of adsorbent layer/column, m 0.6 0.2 0.4
Wall thickness of column, m 0.0024
Internal diameter of column, m 0.016
Density of column wall, kg/m3 8238
Inter-particle voidage/column porosity 0.33
Specific heat capacity, J/kg/K 500 920 880
Bulk density/ column density, kg/m3 756 715
Thermal conductivity of column wall, W/m/K 13.40
Particle/extrude density of zeolite13X 1130 1040
Particle/ extrude radius of zeolite13X, m 0.0008 0.0009
Intra-particle voidage/extrude porosity 0.54 0.46
Extrude tortuosity 2.2 2

65
Table 4.5: Parameters used in simulation of zeolite13X-CMS3K system

Zeolite13X CMS3K
Total Pressure, bars 2.5, 5, 7.5, 10
Flow rate, SLPM 1.6
Temperature, K 300
Mass transfer coefficient, 1/s CH4 – 0.403 CH4 – 0.00004
CO2 – 0.018 CO2 – 0.0086
N2 – 0.112 N2 – 0.0047
Heat transfer coefficient, W/m2/K Gas-Solid – 60 Gas-Solid – 65
Gas-wall – 36 Gas-wall – 39
Wall-ambient – 192 Wall-ambient – 192
Isotherm parameters for CH4 a1 – 0.0112 a1 – 0.000171
a2 – 0.00029 a2 – 0.000172
a3 – 0.834 a3 – 0.716
a4 – 1590 a4 – 3190
a5 – 0.00037 a5 – 0.00001
a6 – 1610 a6 – 2980
Isotherm parameters for CO2 a1 – 0.00033 a1 – 0.00049
a2 – 0.0021 a2 – 0.0105
a3 – 0.43 a3 – 0.492
a4 – 2850 a4 – 2160
a5 – 0.0031 a5 – 0.00147
a6 – 2450 a6 – 2040
Isotherm parameters for N2 a1 – 0.0022 a1 – 0.000003
a2 – 0.00042 a2 – 0.871
a3 – 0.885 a3 – 0.796
a4 – 1740 a4 – 1460
a5 – 0.0001 a5 – 0.0011
a6 – 1970 a6 – 1360
Note: Transport coefficients were calculated (70%CH4/ 20%CO2/10% N2/2.5bars/300K).
66
100 100
90 90

N2 Separation efficiency (%)


N2 Separation efficiency (%)

80 80
70 70
60 60
50 50
40 40
75CH4-15N2-10CO2 75CH4-15N2-10CO2
30 65CH4-15N2-20CO2 30 65CH4-15N2-20CO2
20 65CH4-25N2-10CO2 20 65CH4-25N2-10CO2
55CH4-25N2-20CO2 55CH4-25N2-20CO2
10 10
0 0
0.00 1.00 2.00 3.00 4.00 5.00 0.00 0.50 1.00 1.50 2.00
Total amount of adsorbent Total amount of adsorbent
(kg/mol of feed gas) (kg/mol of feed gas)

(a) 2.5 bars (b) 5.0 bars

100 100
90 90
N2 Separation efficiency (%)
N2 Separation efficiency (%)

80 80
70 70
60 60
50 50
40 40
75CH4-15N2-10CO2 75CH4-15N2-10CO2
30 65CH4-15N2-20CO2 30 65CH4-15N2-20CO2
20 65CH4-25N2-10CO2 20 65CH4-25N2-10CO2
55CH4-25N2-20CO2 55CH4-25N2-20CO2
10 10
0 0
0.00 0.50 1.00 1.50 2.00 0.00 0.20 0.40 0.60 0.80 1.00 1.20
Total amount of adsorbent Total amount of adsorbent
(kg/mol of feed gas) (kg/mol of feed gas)

(c) 7.5 bars (d) 10.0 bars

(Note: 75CH4-15N2-10CO2 means 75 mol % CH4, 15 mol % N2 and 10 mol CO2. Other

legends shall be read same way)

Figure 4.7: Total amount (kg/mol of feed gas) of adsorbents for N2 and CO2 separation

from natural gas for zeolite13X-CMS3K adsorber

67
4.3.1 Parametric study

Parametric studies were performed to determine the effect of feed conditions on

adsorption of CO2 and N2. Three effects were determined. They are: (i) the effect of feed

pressure (ii) the effect of concentration and (iii) the effect of N2 separation efficiency.

Effect of feed pressure: Figure 4.8 plots nitrogen separation efficiency of a

double bed (zeolite13X-CMS3K) adsorption column against the total amount of

adsorbent at various pressures for two different feed compositions: (i) 75% CH4-15% N2-

10% CO2 and (ii) 55% CH4-25% N2-20% CO2. As being evident, feed pressure has

significant effects on the separation of CO2 and N2. High feed pressure requires less

adsorbent, while low feed pressure requires more adsorbent. Also, the effect of the

differential pressure increment, 2.5 bars in this study, on the amount of adsorbent for a

particular separation efficiency became reduced.

Figure 4.9 plots the effect of feed-gas pressure on the required amount of

adsorbent for three separation efficiencies (70%, 80%, and 90%). The distinct points,

when fitted, show power law relationships. As evident from this figure (4.9), a pressure

increase of 2.5 bars from 2.5 to 5 bars for 80% N2 separation is 1.18 kg/mol, for 5 to 7.5

bars, the amount is 0.4 kg/mol, and for 7.5 bars to 10 bars, the amount is 0.24 bar. These

reduced differential amounts also explain the saturation of adsorption capacity of

adsorbents at high pressure.

68
100

N2 separation efficiency (%)


90
80
70
60
50
40
30 Feed gas pressure, 2.5 bar
Feed gas pressure, 5 bar
20
Feed gas pressure, 7.5 bar
10 Feed gas pressure, 10 bar
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Total amount of adsorbent
(kg per mole of feed gas)

(a) 75% CH4-15% N2-10% CO2

100
90
N2 separation efficiency (%)

80
70
60
50
40
Feed gas pressure, 2.5 bar
30
Feed gas pressure, 5 bar
20 Feed gas pressure, 7.5 bar
10 Feed gas pressure, 10 bar
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Total amount of adsorbent
(kg per mole of feed gas)

(b) 55% CH4-25% N2-20% CO2

Figure 4.8: Effect of feed pressure on N2 separation efficiency (%) for

zeolite13X-CMS3K system

69
4.0

3.5
70%
Total amount of adsorbent
(kg per mole of feed gas)

3.0 80%
90%
2.5

2.0

1.5

1.0

0.5

0.0
0.0 2.5 5.0 7.5 10.0
Feed gas pressure (bar)
Note: Total amount denotes amount of zeolite13X and CMS3K.

Figure 4.9: Effect of feed-gas pressure on total amount of adsorbents for 70 to 90% N2

separation efficiency for zeolite13X-CMS3K system

70
Effect of feed composition: Presence of N2 and CO2 in various amounts may

affect the nitrogen separation efficiency of the double bed adsorber. Since all the CO2

was removed in zeolite13X, CO2 has no effect on the N2 separation efficiency of

CMS3K. However, zeolite13X also adsorbs some N2, and the amount of zeolite13X

needed for CO2 separation depends on the concentration of CO2 in the feed. Thus, the

effects of concentration of CO2 may not be significant on overall efficiency, but it will

have a positive effect on the N2 separation efficiency of CMS3K. Overall, the effect of N2

concentration on separation efficiency will be reduced. A further reduced effect was

expected as the double bed adsorber was operated in short cycle time to facilitate kinetic

separation.

Figure 4.10 shows the effect of feed composition on nitrogen separation

efficiency at 2.5 bars and 300K. Feed compositions did not show any significant effect on

the required amount of adsorbent for a specified nitrogen separation efficiency. This is

due to the short cycle time, which limits the ability of the mass or heat transfer parameter,

such as diffusion or conduction, to play a significant role. Nitrogen separation efficiency

was found to vary a maximum of 3.5% (AAD, average absolute deviation) at 2.5 bars.

Since the results of other feed pressures showed a similar trend, they are not shown here.

Figure 4.10a presents a comparison of the effect of the change in concentration of N2.

Here, the concentration of CO2 is constant, while the concentration of CH4 changes. The

efficiency loss, 3.5%, is due to presence of extra N2 in the feed. A similar result was

obtained for 20% CO2. Again, the loss in adsorption efficiency is due to extra N2 that

enters the column.

71
100
90

N2 separation efficiency (%)


80
70
60
50
40
30 65M-15N-20C@2.5bar
55M-25N-20C@2.5bar
20
10
0
0 1 2 3 4 5
Total amount of adsorbent
(kg per mole of feed gas)

(a) Fixed concentration of CO2

100

90
N2 separation efficiency (%)

80

70

60

50

40

30 65M-25N-10C@2.5bar
20 55M-25N-20C@2.5bar

10

0
0 1 2 3 4 5
Total amount of adsorbent
(kg per mole of feed gas)

(b) Fixed concentration of N2

Figure 4.10: Effect of feed concentration on nitrogen separation efficiency at 2.5 bars for

zeolite13X-CMS3K system

72
Effect of N2 separation efficiency: Figure 4.11 plots the total amount of

adsorbent (kg per mole of feed gas) against the N2 separation efficiency of the

zeolite13X-CMS system. As is evident, the there is less of an effect at high feed pressure

(10 bars), while it is significant at low feed-gas pressure (2.5 bar). In these two plots, the

concentration of CO2 was kept constant (10 %), while concentrations of N2 were

changed: 25% N2 (Figure 4.11(a)) and 15% N2 (Figure 4.11(b)). In both cases, the trends

of effects are similar.

4.3.2 Correlations based on simulated results

The effects of feed pressure were correlated to give the amount of total adsorbent

needed for nitrogen separation. The correlation (4.2) shown below reproduces the amount

of total adsorbent at pressures of 2.5, 5, 7.5, and 10 bars with AADs of 3.36%, 4.23%,

3.75%, and 5.28 % for all feed compositions.

b1 ( Pnb 2 )  (nb3 )


Q 4.2
b4  b5  ( Pnb 2 )  (nb3 )

where Q is the total amount (kg per mole of feed gas) of adsorbent, Pn is the partial

pressure of N2, ηn is the separation efficiency (%), and b1 to b5 are constants (Table 4.6).

The results produced by the correlation were compared to simulated results. Figure 4.12

shows the comparison of the results for all pressures and for the feed composition of 75%

CH4, 15% N2, and 10 % CO2. The average absolute deviations observed for this feed

composition were 3.63%, 2.29%, 1.03%, and 3.71%.

73
5

Feed gas pressure, 2.5 bar


4 Feed gas pressure, 5 bar

Total amount of adsorbent


(kg per mole of feed gas)
Feed gas pressure, 7.5 bar
Feed gas pressure, 10 bar
3

0
0 20 40 60 80 100
N2 separation efficiency (%)

(a) 75% CH4-25% N2-10% CO2

Feed gas pressure, 2.5 bar


4 Feed gas pressure, 5 bar
Total amount of adsorbent
(kg per mole of feed gas)

Feed gas pressure, 7.5 bar


Feed gas pressure, 10 bar
3

0
0 20 40 60 80 100
N2 separation efficiency (%)

(b) 55% CH4-15% N2-10% CO2

Figure 4.11: Effect of N2 separation efficiency on total amount of adsorbent for

zeolite13X-CMS3K system

74
Table 4.6: Parameters for Correlation 4.2

Feed Pressure (bars)


parameter
2.5 5.0 7.5 10.0
b1 0.362 0.410 0.401 0.189
b2 0.030 0.011 0.010 0.032
b3 0.625 0.499 0.263 1.040
b4 6.801 10.376 9.576 40.856
b5 -0.311 -0.880 -2.626 -0.176

75
4
Predicted (kg per mole of feed gas)

Feed gas pressure, 2.5 bar


Feed gas pressure, 5 bar
3 Feed gas pressure, 7.5 bar
Feed gas pressure, 10 bar

0
0 1 2 3 4
Simulated (kg per mole of feed gas)

Note: Simulated means results obtained from Aspen Adsorption while


predicted means results obtained from empirical correlation.

Figure 4.12: Comparison of simulated result with the results obtained from correlation

4.2 for feed composition of 75% CH4, 15% N2 and 10 % CO2 for zeolite13X-CMS3K

system

76
4.4 Simulation results for zeolite13X-ACB system

A double bed adsorber consisting of zeolite13X-ACB was used to determine the

amount of adsorbent for complete separation of CO2 and for various N2 separation

efficiencies. The first bed carried zeolite13X, which selectively removed CO2, and the

second bed consisted of ACB, which removed N2. The zeolite13X bed was applied first

to remove all the CO2 from the feed. ACB was then set to remove N2 from remaining gas

mixture. Five pressure stages (2.5 bars, 5 bars, 10 bars, 20 bars, and 30 bars) were

studied using 5 to 25% N2, 5 to 25% CO2, and 70% CH4.

A sample of inputs is tabulated in Table 4.7 and Table 4.8. The parameters shown

in the tables were kept unchanged during simulation. Some other properties, such as mass

transfer coefficients, changed with concentration or pressure changes. The feed was

continued in 10 minutes cycles at a constant flow rate of 1.16 SLPM. This cycle time (10

minutes) was decided on by analyzing several breakthrough simulations with different

ratios of activated carbon to zeolite13X at 2.5 bars with a feed of 70% CH4, 20% CO2,

and 10% N2. The total amount of adsorbents needed for various levels of separation was

determined through simulation and is presented in Figure 4.13. According to Figure 4.13,

the highest amount of adsorbent (916 kg/kmol) was needed for 100% separation of CO2

and 95% of N2 from a feed stream of 70% CH4, 5% CO2, and 25% N2. Most of the

amount was activated carbon. The share of zeolite13X in this amount is only 4.35% (by

weight). Zeolite13X also adsorbed 2.92% of N2. The rest, 92.08 % nitrogen, was

adsorbed in activated carbon.

77
Table 4.7: Physical properties of double bed adsorber (zeolite13X-ACB)

(Cavenati et al. (2006) and Shen et al. (2010))

Column Zeolite13X ACB


Height of adsorbent layer/column, m 0.6 0.2 0.4
Wall thickness of column, m 0.0024
Internal diameter of column, m 0.016
Density of column wall, kg/m3 8238
Inter-particle voidage/column porosity 0.33 0.32
Specific heat capacity, J/kg/K 500 920 650
Bulk density/column density, kg/m3 756 669
Thermal conductivity of column wall, W/m/K 13.40
Particle/extrude density of zeolite13X 1130 984
Particle/extrude radius of zeolite13X, m 0.0008 0.0006
Intra-particle voidage/extrude porosity 0.54 0.51
Tortuosity 2.2 2

78
Table 4.8: Parameters used in simulation of zeolite13X-ACB system

Zeolite13X ACB
Total Pressure, bars 2.5, 5, 10, 20, 30
Flow rate, SLPM 1.6
Temperature, K 300
Mass transfer coefficient, 1/s CH4 – 0.403 CH4 – 0.1252
CO2 – 0.018 CO2 – 0.0963
N2 – 0.112 N2 – 0.7030
Heat transfer coefficient, W/m2/K Gas-Solid – 60 Gas-Solid – 62
Gas-wall – 36 Gas-wall – 39
Wall-ambient – 192 Wall-ambient – 192
Isotherm parameters for CH4 a1 – 0.0112 a1 – 0.00057
a2 – 0.00029 a2 – 0.00059
a3 – 0.834 a3 – 0.8419
a4 – 1590 a4 – 2484
a5 – 0.00037 a5 – 0.00005
a6 – 1610 a6 – 2663
Isotherm parameters for CO2 a1 – 0.00033 a1 – 0.00123
a2 – 0.0021 a2 – 0.00287
a3 – 0.43 a3 – 0.694
a4 – 2850 a4 – 1996
a5 – 0.0031 a5 – 0.00001
a6 – 2450 a6 – 2972
Isotherm parameters for N2 a1 – 0.0022 a1 – 0.0107
a2 – 0.00042 a2 – 0.00035
a3 – 0.885 a3 – 0.8491
a4 – 1740 a4 – 1362
a5 – 0.0001 a5 – 0.000001
a6 – 1970 a6 – 2486
Note: Transport coefficients were calculated (70%CH4/ 20%CO2/10% N2/2.5bars/300K).
79
100 100
90 90

N2 separation efficiency(%)

N2 separation efficiency(%)
80 80
70 70
60 60
50 50
40 40
30 10% CO2 - 20% N2 30 10% CO2 - 20% N2
20 15% CO2 - 15% N2 15% CO2 - 15% N2
20
20% CO2 - 10% N2
10 10 20% CO2 - 10% N2
0 0
0 200 400 600 800 1000 0 100 200 300 400 500
Total amount of adsorbent Total amount of adsorbent
(g/mol of feed gas) (g/mol of feed gas)
(a) 2.5 bars (b) 5 bars
100 100
90 90
N2 separation efficiency(%)

80
N2 separation efficiency(%)
80
70 70
60 60
50 50
40 40
30 10% CO2 - 20% N2 30
10% CO2 - 20% N2
20 15% CO2 - 15% N2 20
15% CO2 - 15% N2
10 20% CO2 - 10% N2 10 20% CO2 - 10% N2
0 0
0 50 100 150 200 250 300 350 0 50 100 150 200
Total amount of adsorbent Total amount of adsorbent
(g/mol of feed gas) (g/mol of feed gas)
(c) 10 bars (d) 20 bars
100
90
N2 separation efficiency(%)

80
70 Figure 4.13: Total amount of
60
50 adsorbents for N2 separation at
40
30 different feed pressures and
10% CO2 - 20% N2
20
15% CO2 - 15% N2
10 compositions (zeolite13X-ACB
20% CO2 - 10% N2
0
0 50 100 150 200 system)
Total amount of adsorbent
(g/mol of feed gas)

(e) 30 bars

80
4.4.1 Parametric study

Parametric studies were performed to determine the effect of feed conditions on

the adsorption of CO2 and N2. Three effects were determined. They are: (i) the effect of

feed pressure (ii) the effect of concentration and (iii) the effect of N2 separation

efficiency.

Effect of feed pressure: Feed pressure has considerable impact on the required

amount of adsorbents. High feed pressure required a low amount of adsorbent, while a

high amount of adsorbent was needed for low pressure. Figure 4.14 shows the amount of

adsorbents required for different pressures for a feed stream of 70% CH4, 15% CO2, and

15% N2. A maximum of 783 g/mol of adsorbents were needed to separate 95% of N2 at

2.5 bars while the number for 30 bars is 156 g/mol. The amount is reduced by 627 g/mol

when pressure was changed to 2.5 bars from 30 bars. A similar trend was also found for

other feed composition and separation levels.

81
800

700
Total amount of adsorbent

95% N2 separation
600
(g/mol of feed gas)

85% N2 separation
500 75% N2 separation

400

300

200

100

0
0 5 10 15 20 25 30
Feed gas pressure (bar)

Figure 4.14: Effect of feed pressure on total amount of adsorbent for different N2

separation efficiencies (zeolite13X-ACB system)

82
Effect of feed composition: Figure 4.15 shows the total amount of adsorbent as

a function of concentration of CO2 and N2 at various feed pressures. The top plot of the

figure shows that the total amount of adsorbent decreases with an increasing

concentration of CO2, while the bottom plot shows an increase in adsorbent amount. The

conflict reveals that the amount of activated carbon for N2 removal is higher when

compared to zeolite13X for CO2 removal. This can be credited to the equilibrium

adsorption capacity of zeolite13X and ACB. Another notable point is the diminishing

nature of concentration effect at higher pressure. This indicates that the adsorption

capacity of adsorbents becomes limited after a certain pressure.

Effect of N2 separation efficiency: Figure 4.16 plots the total amount of required

adsorbents as a function of N2 separation efficiency. The plots were drawn for the lowest

feed-gas pressure (2.5 bars) and highest feed-gas pressure (30 bars) of this study. The

feed contained 25%, 15%, and 5% N2. The other two concentrations (10% N2 and 20%

N2) were dropped for clarity of the plots. As evident from the plots, the amount of

adsorbent is high for high separation efficiency. The relationship seems linear at low feed

pressure, while at high pressure, it is not. Also, the relationship changes for over 90%

separations. The starting point of separation efficiency refers to adsorbed amount of N2 in

zeolite13X.

83
1000
900

Total amount of adsorbents


800
Feed gas at 2.5 bar
700 Feed gas at 5 bar

(g/mol of feed gas)


Feed gas at 10 bar
600 Feed gas at 20 bar
Feed gas at 30 bar
500
400
300
200
100
0
0 5 10 15 20 25
CO2 concentration (%)

(a) Effect of concentration of CO2

1000
900
Total amount of adsorbents

800
Feed gas at 2.5 bar
700 Feed gas at 5 bar
(g/mol of feed gas)

Feed gas at 10 bar


600 Feed gas at 20 bar
Feed gas at 30 bar
500
400
300
200
100
0
0 5 10 15 20 25 30
N2 concentration (%)

(b) Effect of concentration of N2

Figure 4.15: Effect of concentration on total amount of adsorbents at different feed

pressures for zeolite13X-ACB system

84
1000
900 25% N2
800

Total amount of adsorbent


15% N2
700 5% N2

(g/mol of feed gas)


600
500

400
300

200
100

0
0 10 20 30 40 50 60 70 80 90 100
N2 separation efficiency (%)

(a) Feed gas pressure 2.5 bars

200
180 25% N2
160
Total amount of adsorbent

15% N2
140 5% N2
(g/mol of feed gas)

120
100

80
60

40
20

0
0 10 20 30 40 50 60 70 80 90 100
N2 separation efficiency (%)

(b) Feed gas pressure, 30 bars

Figure 4.16: Effect of N2 separation efficiency on total amount of adsorbent for feed

pressures of (a) 2.5 bars and (b) 30 bars (zeolite13X-ACB system)

85
4.4.2 Correlations based on simulated results

The amounts of adsorbents determined from the feed pressures were correlated in

a single equation:

d1 S d 2  (d 3  d 4  P d 5 )  y d 6  (1  y) d 7
Q 4.3
d 8  d 9  P d 5  y d 10  (1  y) d 11

where w is the amount of adsorbent, S is the separation factor, P is the feed pressure, y is

the mole fraction of N2 or CO2, and d1 to d11 are parameters. The correlation determines

the total amount of adsorbents and the amount of zeolite13X depending on the input

parameters shown in Table 4.7. There are 2 sets of parameters listed in this table. They

entail the total amount of required adsorbent depending on the N2 content of the feed.

Figure 4.20 shows the amount of adsorbents obtained from simulation (points) and the

prediction of the correlation for the feed stream of 70% CH4, 25% CO2, and 05% N2. The

predicted results deviated from the simulated results by 3.8% (AAD). Maximum

deviation was observed for 85% nitrogen separation at 5 bars. The amount of adsorbent

under this condition was under the amount predicted by 4.6%. The predicted results for

the feed stream that carried 70% CH4, 05% CO2, and 25 % N2 showed an average

absolute deviation of 1.8%. Maximum absolute deviation was 5.5%.

86
Table 4.9: Parameters for correlation 4.3

Nitrogen (mol/mol)
Parameter
y<0.15 y≥0.15
d1 4.773 3.719
d2 1.026 1.032
d3 1.648 1.315
d4 0.413 0.241
d5 0.039 0.108
d6 0.042 0.070
d7 0.516 0.118
d8 32.542 84.280
d9 0.079 -1.411
d10 1.502 -1.182
d11 0.912 2.653

87
1000

95% N2 separation
Predicted amount (gm/mol)

800 85% N2 separation


75% N2 separation

600

400

200

0
0 200 400 600 800 1000
Simulated amount (gm/mol)

Note: Simulated means results obtained from Aspen Adsorption while


predicted means results obtained from empirical correlation.

Figure 4.17: Comparison of simulated and predicted (correlation 4.3) results

88
4.5 Determination of column dimensions using correlations

The procedure for the determination of column dimensions is described below. A

flow diagram, Figure 4.18, is, also, included.

Step1- Determination of amount of zeolite13X: Using correlation 4.1, the

required amount of zeolite13X for complete separation of CO2 can be determined. The

required inputs are feed pressure and concentration (mole fraction) of CO2. Parameters

for the correlation can be found in Table 4.3.

Step2 - Determination of total amount of adsorbent: The total amount of

adsorbent can be determined from correlations in Tables 4.2 or 4.3 for zeolite13X-

CMS3K and zeolite13X-ACB systems, respectively. The required inputs are feed

pressure, concentration of nitrogen, and separation efficiency (%) or separation factor for

zeolite13X-CMS3K and zeolite13X-ACB systems, respectively.

Step3 - Determination of amount of CMS3K and ACB: The amount of CMS3K

or ACB is the difference of the amounts obtained using correlations from Tables 4.2 and

4.1 or 4.3 and 4.1, respectively.

Step4 - Determination of volume of adsorbents: The total volume of adsorbents

is the sum of the volume of zeolite13X (determined at Step-1) and the volume of CMS3K

or ACB. The volume of individual adsorbents can be obtained by dividing the adsorbent

amount by its bulk density. Bulk density can be found in Tables 4.1, 4.4, or 4.7 for

zeolite13X, CMS3K, and ACB, respectively.

Step5 - Determination of dimensions of column: Flow velocity of the pipeline

can be converted into instantaneous velocity by using column porosity. The obtained

velocity is then lowered to accommodate retention time. This lowered velocity can be

89
used to obtain the cross-sectional area of the column. Then, column length can be

determined by dividing the total volume of adsorbent by cross-sectional area of the

column.

Note: Diameter does not affect the performance of the column. The system can be used

for larger diameter.

90
Process conditions

Pressure Pressure Flow rate


Concentration of CO 2 Concentration of CO 2 and N2
Required N2 separation efficiency

Correlation 4.1 Correlation 4.2 or 4.3 Column


Amount (g/mol) Total amount (g/mol) diameter
Zeolite13X Zeolite13X+CMS3K/ACB

Porosity
zeolite13X Amount (g/mol)
CMS3K or ACB

Prosity
CMS3K/ACB

Volume Volume Volume Column


Zeolite13X Total CMS3K or ACB Length

Figure 4.18: Calculation procedure for determination of column dimension using

correlations

91
5 Conclusions and Recommendations for future work

5.1 Conclusions

We developed a two-step design method for a layered bed adsorber for the

selective separation of CO2 and N2 from natural gas using zeolite13X, CMS3K, and

ACB. This two-step design method determines important design parameters and the

amount of adsorbents and then transforms the amount of adsorbents into physical

dimensions of the column. This method will be useful in designing practical separation

processes to produce pipeline grade NG.

Mathematical correlations, the first of their kind, were developed as a part of this

design process. These correlations determine the amount of adsorbents using information

such as feed pressure, concentration of CO2 and N2, and the extent of desired separation.

Moreover, these correlations show the way to derive preliminary designs at reduced cost

as they eliminate extensive simulation or experiments.

A step by step procedure was outlined for transforming the information obtained

by correlation into physical dimensions of the column. The procedure also offers a

flexible opening for gas velocity. Usual pipeline velocity can be used to obtain design

parameters. The flexibility in velocity will help the designer to keep control of column

dimensions such as diameter and length.

A parametric study was performed on a single-bed adsorption system to evaluate

(i) the contribution of each of the three mass transfer resistances present in the adsorption

separation process and (ii) the contribution of heat transfer modes. The study revealed

that (i) macropore resistance was dominant in zeolite13X, and (ii) convective heat

transfer was more pronounced than the conductive heat transfer in gas and adsorbent.
92
A parametric study was performed on a layered bed (zeolite13X-CMS3K and

zeolite13X-ACB) adsorption system. One notable finding was the reduction of CO2

adsorption capacity of zeolite13X at high pressure (approximately 7.5 bars and over) in

the presence of N2. Another noteworthy observation was the reversal of selectivity of

CO2 over N2 on zeolite13X. In zeolite13X-CMS3K systems, the effects of N2 on

concentration were found to be insignificant, while zeolite13X-ACB systems showed

considerable effects.

Parameters of the isotherm model were obtained from the fitting of experimental

data. Several isotherm models were tried. The temperature dependency of those isotherm

models were also analyzed to account for temperature variations due to the heat of

adsorption. The temperature-dependent Langmuir-Freundlich model was found to be the

best fit.

93
5.2 Recommendations for future work

In this work, multicomponent adsorption was implemented with the help of the

ideal adsorption solution theory in the absence of experimental data on multicomponent

adsorption. The theory uses pure component equilibrium information to calculate the

mixture properties. Experimental study of multicomponent gas mixture shall be done.

This will help to determine the optimum design pressure of a double bed adsorber.

94
References

Alvarez-Ramırez, J., Fernandez-Anaya, G., Valdes-Parada, F.J. and Ochoa-Tapia, J.A.

(2005). Physical consistency of generalized linear driving force models for

adsorption in a particle. Industrial and Engineering Chemistry Research, 44 (17),

6776-6783.

Barrer, R. M., Ash, R., and Pope, C. G. (1963). Flow of adsorbable gases and vapours in

a microporous medium I. Single sorbates. Proceedings of the Royal Society of

London, A271, 1-18, 19-31

Bastos-Neto, M., Moeller, A., Bohm, J., Glaser, R. and Staudt, R. (2011). Breakthrough

Curves of Methane at High Pressures for H2 Purification Processes. Chemie

Ingenieur Technik, 83, 183-190.

Biegler, L.T., Jiang, L. and Fox, V.G. (2004). Recent advances in simulation and optimal

design of pressure swing adsorption systems. Separation and Purification

Reviews, 33, 1–39

Bird, R. B., Stewart, W. E., and Lightfoot, E. N. (1960). Transport phenomena. John

Wiley and Sons Ltd, New York

Book, D. L. (1981). Finite-Difference Techniques for Vectorized Fluid Dynamics

Calculations. Springer-Verlag

Cavenati, S., Grande, C.A., and Rodrigues, A.E. (2004). Adsorption equilibrium of

methane, carbon dioxide and nitrogen on zeolite 13X at high pressures. The

Journal of Chemical and Engineering Data, 49: 1095–1101.

95
Cavenati, S., Grande, C.A., and Rodrigues, A.E. (2005). Separation of methane and

nitrogen by adsorption on carbon molecular sieve. Separation Science and

Technology, 40, 2721–2743

Cavenati, S., Grande, C.A., and Rodrigues, A.E. (2006). Separation ofCH4/CO2/N2

mixtures by layered pressure swing adsorption for upgrade of natural gas.

Chemical Engineering Science, 61, 3893 – 3906

Chihara, K., and Suzuki, M. (1983). Simulation of nonisothermal pressure swing

adsorption. Journal of Chemical Engineering Japan, 16, 53-60.

Chlendi, M., Tondeur, D. and Rolland, F. (1995). Method to obtain a compact

representation of process performances from a numerical simulator: Example of

pressure swing adsorption for pure hydrogen production. Gas Separation and

Purification, 9, 125–135

Chlendi, M., Tondeur, D. and Rolland, F. (1995). A method to obtain a compact

representation of process performances from a numerical simulator: example of

pressure swing adsorption for pure hydrogen production. Gas Separation and

Purification, 9, 125-135

Choi, S. Drese, J. H. and Jones, C. W. (2009). Adsorbent materials for carbon dioxide

capture from large anthropogenic point sources. ChemSusChem, 796 - 854

Cruz, P., Santos, J. C., Magalhaes, F. D., and Mendes, A. (2003). Cyclic adsorption

separation processes: analysis strategy and optimization procedure. Chemical

Engineering Science, 58, 3143-3158.

Da Silva, F.A., Silva, J.A., and Rodrigues, A.E. (1999). A general package for the

simulation of cyclic adsorption processes. Adsorption, 5, 224–229.

96
Ding, Y. and Alpay, E. (2000). Equilibria and kinetics of CO2 adsorption on hydrotalcite

adsorbent. Chemical Engineering Science, 55, 3461-3474.

Do, D. D. (1998). Adsorption analysis: Equilibria and Kinetics. Imperial College Press,

London

Doong, S. J., and Yang, R. T. (1986). Parametric study of the pressure swing adsorption

process for gas separation: A criterion for pore diffusion limitations. Chemical

Engineering Communication, 41, 163-180.

EIA, (1999). Natural gas 1998, issues and trends, DOE/EIA-0560(98), U.S. Department

of Energy

Farooq, S., Hassan, M. M., and Ruthven, D. M. (1988). Heat effects in pressure swing

adsorption systems. Chemical Engineering Science, 43, 1017-1031.

Farooq, S. and Ruthven, D.M. (1991). Numerical simulation of kinetically controlled

pressure swing adsorption bulk separation based on a diffusion model. Chemical

Engineering Science, 46, 2213-2224

Farooq, S. and Ruthven, D. M., 1990, A comparison of linear driving force and pore

diffusion models for a pressure swing adsorption bulk separation process.

Chemical Engineering Science, 45, 107-l 15.

Farooq, S., Qinglin, H., and Karimi, I.A. (2002). Identification of transport mechanism in

adsorbent micropores from column dynamics. Industrial and Engineering

Chemistry Research, 41, 1098-1106.

Finlayson, B. A. (1992). Numerical methods for problems with moving fronts. Ravenna

Park Publishing Inc, USA

97
Flores-Fernandez, G. and Kenney, C.N. (1983). Modeling of the pressure swing air

separation process. Chemical Engineering Science, 38, 827-834

Foley, H.C. (1995). Carbogenic molecular-sieves—Synthesis, properties and

applications, Microporous Mesoporous Mater. 4, 407–433

Fuller, E. N., Schettler, P. D., and Giddings, J. C. (1966). A new method for prediction

of binary gas-phase diffusion coefficients. Industrial and Engineering Chemistry,

58, 19-27.

Gear, C. W. (1971). Numerical initial value problems in ordinary differential equations.

Prentice-Hall, Englewood Cliffs, NJ.

Glueckauf, E. (1949). Theory of chromatography VII. The general theory of two solutes

following non-linear isotherms. Discussions of the Faraday Society, 7, 12-25

Glueckauf, E. (1955). Theory of chromatography part 10 – Formula for diffusion into

spheres and their application to chromatography. Transactions of the Faraday

Society, 51, 1540-1551.

Grisky, R.G., 2002. Transport Phenomena and Unit Operations. John Wiley and Sons

Ltd., New York.

Habgood, H.W. (1958). The kinetics of molecular sieve action. Sorption of nitrogen,

methane mixtures by Linde Molecular Sieve 4A. Canadian Journal of Chemistry,

36, 1384-1397.

Harlick, P. J. E. and Tezel, F. H. (2004). An experimental adsorbent screening study for

CO2 removal from N2. Microporous Mesoporous Mater, 76, 71–79

Hassan, M.M. (1985). Theoretical and experimental studies of the pressure swing

adsorption system. Ph.D. Thesis, University of New Brunswick, NB, Canada

98
Hassan, M. M., Ruthven, D. M., and Raghavan, N. S. (1986). Air Separation by pressure

swing adsorption on a carbon molecular Sieve. Chemical Engineering Science,

41, 1333-1343

Hirsch, C. (2007). Numerical Computation of Internal and External Flow. John Wiley and

Sons Ltd, New York.

Hobson, G. D., and Tiratso, E. N. (1985). Introduction to petroleum geology, Gulf

Publishing, Houston, TX.

Haynes, H. W. Jr. (1975). The determination of effective diffusivity by gas

chromatography. Time domain solutions. Chemical Engineering Science, 30, 955-

961

Jackson, R. (1977). Transport in porous catalysts. Elsevier, New York

Jee, J. G., Kim, M. B. and Lee, C. H. (2001). Adsorption Characteristics of Hydrogen

Mixtures in a Layered Bed: Binary, Ternary and Five-Component Mixtures.

Industrial and Engineering Chemistry Research, 40, 868-878.

Jiang, L., Biegler, L., Fox, V. (2003). Simulation and optimization of pressure swing

adsorption systems for air separation. AIChE Journal, 49(5), 1140-1157

Kidnay, A. J. and Parrish, W. R.. (2006). Fundamentals of natural gas processing. Taylor

and Francis: Boca Raton.

Kikkinides, E. S., and Yang, R. T. (1993). Gas separation and purification by polymeric

adsorbents: Flue gas desulfurization and SO2 recovery with styrenic polymers.

Industrial and Engineering Chemistry Research, 32, 2365-2372.

99
Ko, D., Siriwardane, R. and Biegler, L. T. (2003). Optimization of pressure swing

adsorption process using zeolite 13X for CO2 sequestration. Industrial and

Engineering Chemistry Research, 42, 339-348.

Krishnan, S. V. (1993). Ph.D. Dissertation, University of Rochester, Rochester, NY.

Kumar, R. (1990). Adsorptive process for producing two gas streams from a gas mixture.

US Patent 4,915,711 assigned to Air Products and Chemicals.

Kumar, R., Fox, V.G., Hartzog, D.G., Larson, R.E., Chen, Y.C., Houghton, P.A., and

Naheiri, T. (1994). A versatile process simulator for adsorptive separations.

Chemical Engineering Science, 49, 3115-3125.

Leavitt, F. W. (1962). Non-isothermal adsorption in large fixed beds. Chemical

Engineering Progress, 58, 54-59.

Lee, C. H., Yang, J., and Ahn, H. (1999). Effects of Carbon-to-Zeolite Ratio on Layered

Bed H2 PSA for Coke Oven Gas. AIChE Journal, 45, 535-545.

Lapidus, L. and Amundson, N. R. (1952). Mathematics of adsorption in beds. VI. The

effect of longitudinal diffusion in ion exchange and chromatographic columns.

Journal of Physical Chemistry, 56, 984-988.

Liapis, A. I. and Crosser, O. K. (1982). Comparison of model predictions with

nonisothermal sorption data for ethane-carbon dioxide mixtures in beds of 5A

molecular sieves. Chemical Engineering Science, 37, 958-961

Lu, Z. P., Loureiro, J. M., Levan, M. D., and Rodrigues, A. E. (1992). Effect of

intraparticle forced convection on gas desorption from fixed beds containing

large-pore adsorbents. Industrial and Engineering Chemistry Research, 31, 1530-

1540.

100
Meyer, O. A., and Weber, T. W. (1967). Nonisothermal adsorption in fixed beds. AIChE

Journal, 13, 457-465.

Mitchell, J. E. and Shendalman. L. H. (1972). Study of heatless adsorption in the model

system CO2 in He: part I. Chemical Engineering Science, 27, 1449- 1458

Mitchell, J. E. and Shendalman. L. H. (1973). Study of heatless adsorption in the model

system CO2 in He: part II. AIChE Symposium Series, 69, 25.

Nagel, G., Kluge, G., and Flock, W. (1987). Modeling of non-isothermal multicomponent

adsorption in adiabatic fixed beds - I. The numerical solution of the parallel

diffusion model. Chemical Engineering Science, 42, 143-153.

Nakao, S. and Suzuki, M. (1983). Mass transfer coefficient in cyclic adsorption and

desorption. Journal of Chemical Engineering of Japan, 16, 114-119.

Nilchan, S. (1997). On the optimization of periodic adsorption processes, Ph.D. thesis,

University of London

Nilchan, S. and Pantelides, C. (1998). On the optimization of periodic adsorption

processes. Adsorption, 4, 113-147.

Mason, E. A., and Malinauskas, A. P. (1983). Gas transport in porous media: The dusty-

gas model. Elsevier, New York.

Myers, A.L. and Prausnitz, J. M. (1965). Thermodynamics of mixed gas adsorption.

AIChE Journal, 11, 121–127.

Randall, J. L. (1992). Numerical Methods for Conservation Laws (13 Godunov’s

Method). Springer Science and Business Media

101
Ribeiro, A.M., Grande C.A., Lopes F.V.S., Loureiro, J.M., and Rodrigues A.E. (2008). A

parametric study of layered bed PSA for hydrogen purification. Chemical

Engineering Science, 63, 5258–73.

Rigby, S.P., Fletcher, R.S., and Riley, S.N. (2004). Characterization of porous solids

using integrated nitrogen sorption and mercury porosimetry. Chemical Engineering

Science 59, 41-51

Robertson, E.P. (2007). Analysis of CO2 separation from flue gas, pipeline transportation,

and sequestration in coal. Idaho National Laboratory, Idaho Falls, Idaho

Rosen, J. B. (1952). Kinetics of a Fixed Bed System for Solid Diffusion into Spherical

Particles. Journal of Chemical Physics, 20, 387-397

Rouquerol, F., Rouquerol, I., and Sing, K. (1999). Adsorption by Powders and Porous

Solids-Principles Methodology and Applications, Academic Press, London.

Ruthven, D. M. (2000). Past progress and future challenges in adsorption research

Industrial and Engineering Chemistry Research, 39, 2127-2131

Serbezov, A. and Sotirchos, S. V. (1996). Generalized linear driving force approximation

for multicomponent pressure swing adsorption simulations. AIChE Annual

Meeting, Chicago.

Serbezov, A. and Sotirchos, S. V. (1997a). Mathematical modeling of the adsorptive

separation of multicomponent gaseous mixtures. Chemical Engineering Science,

52, 79-91.

Serbezov, A., and Sotirchos, S. V. (1997b). Multicomponent transport effects in sorbent

particles under pressure swing conditions. Industrial and Engineering Chemistry

Research, 36, 3002-3012.

102
Serbezov, A. and Sotirchos, S. V. (1998). Mathematical modeling of multicomponent

nonisothermal adsorption in sorbent particles under pressure swing conditions.

Adsorption, 4, 93-111.

Serbezov, A. and Sotirchos, S. V. (1999). Particle-bed model for multicomponent

adsorption-based separations: application to pressure swing adsorption. Chemical

Engineering Science, 54, 564-5666

Shen, C., Grande, C.A., Li, P., Yu, J., and Rodrigues, A.E. (2010). Adsorption equilibria

and kinetics of CO2 and N2 on activated carbon beads. Chemical Engineering

Journal,160, 398-407

Shin, H. S. and Knaebel, K. S. (1987). Pressure swing adsorption: A theoretical study of

diffusion-induced separations. AIChE. Journal, 33, 654-662.

Sircar, S. (1979). Separation of Multicomponent Gas Mixtures. US Patent 4,171,206.

Sircar, S., Mohr, R., Ristic, C., and Rao, M. B. (1999). Isosteric Heat of Adosrptions

Theory and Experiment. Journal of Physical Chemistry, 103, 6539-6546.

Sircar, S. and Hufton, J.R. (2000). Why does the linear driving force model for

adsorption kinetics work? Adsorption, 6, 137-147

Siriwardane, R. V., Shen, M. S., and Fisher, E. P. (2005). Adsorption of CO2 on Zeolites

at moderate temperatures. Energy and Fuels, 19, 1153-1159

Smith, J.M. (1970). Chemical Engineering Kinetics, McGraw-Hill, New York.

Sotirchos, S. V. (1989). Multicomponent di!usion and convection in capillary structures.

AIChE Journal, 35, 1953-1961.

Sun, L. M., Le Quere, P., and Levan, M. D. (1996). Numerical simulation of diffusion-

limited PSA process models by finite difference methods. Chemical Engineering

Science, 51, 5341-5352.


103
Takamura, Y., Narita, S., Aoki, J., Hironaka, S. and Uchida,S. (2001). Evaluation of

Dual- Bed Pressure Swing Adsorption for CO2 Recovery from Boiler Exhaust

Gas Separation and Purification and Separations Technology, 24, 519-528.

Tagliabue, M., Farrusseng, D., Valencia, S., Aguado, S., Ravon, U., Rizzo, C., Corma,

A., and Mirodatos, C. (2009). Natural gas treating by selective adsorption:

material science and chemical engineering interplay. Chemical Engineering

Journal, 155, 553-566.

Thomas, H.C., (1944). Heterogeneous ion exchange in a flowing system. Journal of the

American Chemical Society, 66, 1664-1666

Wakao, N. and Funazkri, T. (1978). Effect of fluid dispersion coefficients on particle-to-

fluid mass transfer coefficients in packed beds. Chemical Engineering Science,

33, 1375–1384.

Ward, J. W., and Habgood, H. W. (1966). The infrared spectra of carbon dioxide

adsorbed on Zeolite X. Journal of Physical Chemistry, 70, 1178-1182.

Webley, P. A. and He, J. (2000). Fast solution-adaptive finite volume method for

PSA/VSA cycle simulation; 1 single step simulation. Computers and Chemical

Engineering, 23, 1701-1712

World Energy Council, (2010). 2010 Survey of Energy Resources. ISBN 978 0 946121

021

Yang, J. and Lee, C.H. (1998). Adsorption dynamics of a Layered Bed PSA for H2

recovery from Coke Oven Gas. AIChE Journal, 44, 1325-1334

Yang, R.T. and Doong, S.J. (1985). Gas separation by pressure swing adsorption: a pore

diffusion model for bulk separations. . AIChE Journal, 31, 1829–1842

104
Yang, R. T. (1987). Gas Separation by adsorption process, Butterworth, Boston.

Yang, R. T. (2003). Adsorbents: Fundamentals and Applications, John Wiley and Sons,

Hoboken.

Younger, A. H. (2004). Natural Gas Processing Principles and Technology. University of

Calgary, Canada.

105
Appendix – A: Adjustments of transport parameters

Effects of transport parameters were evaluated for the adsorption of CO2 in

zeolite13X and N2 in ACB and CMS3K. The determined effects helped to maintain a

simplified model without sacrificing the rigour of the model. As an example, highlights

of CO2-zeolite13X systems are described below. The effects were determined using a gas

flow rate 8.74E-07kmol/s (70%CH4, 20% CO2, and 10% N2) in a cylindrical layer of

zeolite (0.20m length x 0.016m diameter) at 2.5 bars and 300K, exactly the same as the

experimental setup previously reported (Cavenati et al., 2005).

Effect of mass transport parameters: The mass transfer analysis was based on the

breakthrough behavior of a fixed-bed adsorption column. The breakthrough curve was

produced using different mass transfer resistances. The resistances were also combined to

check their combined effect on the breakthrough curve. The objective was to identify an

effective mass transfer coefficient for every component of the gas mixture. Three forms

of mass transfer resistances were tested against observed dynamics of the fixed-bed

adsorption process (Figure 4.2). The macropore resistance was identified as a major

contributor for the natural gas-zeolte13X system based on predicted breakthrough

behaviour, i.e., the dynamics of the fixed-bed adsorption column. The observed

dynamics lead to the use of a resistance that is lower than the calculated macropore

resistance. This new value predicted the breakthrough behavior of the column with better

accuracy. It can be concluded that major mass transfer resistances in the zeolite bed exist

in the macropores, and the model takes on simplified forms as two other contributors

becomes less significant and, hence, the computation time will be reduced.

106
0.20 0.05
0.18

CO2 (mol/mol) at bed exit


CO2 (mol/mol) at bed exit

Cavenati et al., 2006


0.16 0.04
Film
0.14
macropore
0.12 0.03 micropore
0.10
0.08 0.02
Cavenati et al., 2006
0.06 Film
0.04 macropore 0.01
micropore
0.02
0.00 0.00
0 400 800 1200 1600 2000 250 300 350 400
Time (second) Time (second)

(a) Effect of single mass transfer resistance

0.20 0.05
Cavenati et al.,
0.18 2006
CO2 (mol/mol) at bed exit

Cavenati et al., Macropore


0.16 0.04
2006
CO2 (mol/mol) at bed exit

0.14
Macropore Film+
0.12 0.03 macropore

0.10 Film+ macropore+


macropore micropore
0.08 0.02
macropore+ Film +
0.06 macropore +
micropore
micropore
0.04 0.01
Film +
0.02 macropore +
micropore
0.00 0
0 400 800 1200 1600 2000 2400 300 320 340 360 380 400
Time (second) Time (Second)

(b) Effect of combined mass transfer resistance

Figure A.1: Breakthrough of CO2 in zeolite13X for various mass transfer resistances

107
Figure A.1(a) shows the individual effect of film, macropore, and micropore

resistances. When compared to the literature breakthrough point (340 seconds), it shows

an extended breakthrough point (370 seconds) for film and micropore resistance and a

short break through (330 seconds) for macropore resistance. This indicates that actual

resistances are larger than macropore resistance and, hence, a combination of resistances

may present in the system. Figure A.1(b) shows the effect of combined resistances.

Again, it is evident that macropore resistance is dominant, though it does not explain the

experimental breakthrough in full. Further investigation was carried out by changing the

bed porosity (Figure A.2). It was found that the bed porosity is related to the macropore

resistance by a factor of ε//(1-ε) rather than ε. Figure A.2 shows the breakthrough

comparison for this modified case. It is noticeable that this result gives better agreement

with literature. Since it is a single case study, we cannot generalize this finding at this

point. However, it certainly can be used in specific cases described in this study. This

modified resistance model was used to analyze the heat transfer issues associated with the

system.

108
0.20
CO2 (mol/mol) at bed exit

0.16

0.12
Cavenati et al., 2006

0.08 Modified Macropore


resistance

0.04

0.00
0 400 800 1200 1600 2000
Time (Second)

Figure A.2: Breakthrough of CO2 in zeolite13X with modified macropore resistance

109
Effect of heat transport parameters: Adsorption is an exothermic process that

releases heat due to fluctuations in the surface energy of solid and thermal energy of

adsorbate molecules. This released heat is partly adsorbed by the solid and experiences a

rise in temperature, which slows down the kinetics and then the dissipation of heat to the

surroundings cools down the solid to facilitate additional adsorption. Hence, knowledge

of heat exchange is critical as it has an influence on local equilibrium and kinetics that

eventually straighten out the separation efficiency. Effects of heat transport parameters

were determined for the same system used for mass transport parameters, keeping mass

transport parameters as constant. It has been found that the convection heat transfer is

dominant in the system. The gas-solid heat transfer coefficient, i.e., the film heat transfer

coefficient depends on local conditions as the variable form of this coefficient diminishes

the temperature gaps between solid and gas phases. As expected, the solid phase and gas

phase conductivity were found to have negligible effects (Figure A.3). Wall conductivity

and external heat transfer are significant in determining shape of the dynamic profiles.

110
0.20

CO2 (mol/mol) at bed exit


0.16

0.12

0.08

No conduction
0.04 gas+solid conduction

0.00
0 400 800 1200 1600 2000
Time (second)

(a) Breakthrough profile

330

325 No
Temperature (K) at bed exit

Conduction
320
gas +
solid phase
315 conduction

310

305

300

295
0 400 800 1200 1600 2000
Time (second)

(b) Temperature profile

Figure A.3: Effect of conductivity (gas and solid) on breakthrough dynamics

111

You might also like