Sunteți pe pagina 1din 32

The removal of CO

2
and N
2
from natural gas: A review of conventional and
emerging process technologies
T.E. Rufford
a
, S. Smart
b
, G.C.Y. Watson
a
, B.F. Graham
a
, J. Boxall
a
, J.C. Diniz da Costa
b
, E.F. May
a,n
a
The University of Western Australia, Centre for Energy, 35 Stirling Highway, Crawley, WA 6009, Australia
b
The University of Queensland, FIMLabFilms and Inorganic Membrane Laboratory, School of Chemical Engineering, College Road, St Lucia, Qld. 4072, Australia
a r t i c l e i n f o
Article history:
Received 5 September 2011
Accepted 2 June 2012
Available online 19 June 2012
Keywords:
CO
2
capture
nitrogen rejection
contaminated gas
membrane
absorption
pressure swing adsorption
hydrates for gas separation
a b s t r a c t
This article provides an overview of conventional and developing gas processing technologies for CO
2
and N
2
removal from natural gas. We consider process technologies based on absorption, distillation,
adsorption, membrane separation and hydrates. For each technology, we describe the fundamental
separation mechanisms involved and the commonly applied process ow schemes designed to produce
pipeline quality gas (typically 2% CO
2
, o3% N
2
) and gas to feed a cryogenic gas plant (typically 50 ppmv
CO
2
, 1% N
2
). Amine absorption technologies for CO
2
and H
2
S removal (acid gas treating) are well-
established in the natural gas industry. The advantages and disadvantages of the conventional amine-
and physical-solvent-based processes for acid gas treating are discussed. The use of CO
2
selective
membrane technologies for bulk separation of CO
2
is increasing in the natural gas industry. Novel low-
temperature CO
2
removal technologies such as ExxonMobils Controlled Freeze Zone
TM
process and
rapid cycle pressure swing adsorption processes are also emerging as alternatives to amine scrubbers in
certain applications such as for processing high CO
2
concentration gases and for developing remote gas
elds. Cryogenic distillation remains the leading N
2
rejection technology for large scale (feed rates
greater than 15 MMscfd) natural gas and liqueed natural gas plants. However, technologies based on
CH
4
selective absorption and adsorption, as well as N
2
selective pressure swing adsorption technol-
ogies, are commercially available for smaller scale gas processing facilities. The review discusses the
scope for the development of better performing CO
2
selective membranes, N
2
selective solvents and N
2
selective adsorbents to both improve separation power and the durability of the materials used in novel
gas processing technologies.
& 2012 Elsevier B.V. All rights reserved.
1. Introduction
In 2010 natural gas (NG) supplied 23.81% of the worlds energy
demand and the volume of natural gas consumed increased by 7.4%
over 2009 levels driven by economic recovery the USA and the
continuing economic development of China, India and Russia (BP
Statistical Review of World Energy June 2011, 2011). The growth in
the global demand for natural gas has led to a re-evaluation of the
development potential of unconventional, stranded and contami-
nated gas reserves that were previously considered economically
unviable. Furthermore, many of the signicant natural gas reserves
are located far from the large established gas markets in Western
Europe, Japan and South Korea. Therefore, signicant volumes of
natural gas must be transported long distances from exporting
countries either by pipeline or by tanker as liqueed natural gas
(LNG); the economics of this choice have been discussed by many
authors such as Rojey et al. (1997). Table 1 contains a list of the
major natural gas importing and exporting countries in 2010
including a breakdown of the amounts transported by pipeline
and LNG. The production of LNG is clearly already essential to the
international trade of natural gas and its importance is set to
increase further over the next two decades, particularly in the
Asia-Pacic region.
The estimated world volumes of sub-quality natural gas
reserves, including sour natural gas reserves, are signicant.
Sub-quality natural gas reserves are dened as gas elds contain-
ing more than 2% CO
2
, 4% N
2
and 4 parts per million (ppm)
hydrogen sulphide (H
2
S) (Kidnay and Parrish, 2006). Burgers et al.
(2011), for example, estimate that 50% of the volume of known
gas resources contain more than 2% CO
2
. The development of
sub-quality, unconventional and remote natural gas reserves,
including development via LNG production, can present new
challenges to gas processing that require more efcient
approaches to the conventional absorption and cryogenic con-
densation technologies that are most commonly used for the
removal of carbon dioxide and nitrogen, respectively, from
Contents lists available at SciVerse ScienceDirect
journal homepage: www.elsevier.com/locate/petrol
Journal of Petroleum Science and Engineering
0920-4105/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.petrol.2012.06.016
n
Corresponding author.
E-mail address: eric.may@uwa.edu.au (E.F. May).
Journal of Petroleum Science and Engineering 94-95 (2012) 123154
natural gas. Additional gas processing challenges arise from new
environmental regulations that may call for capture and seques-
tration of CO
2
from gas elds and stricter regulation of CH
4
emissions in N
2
vent streams from natural gas production
facilities.
Carbon dioxide, as well as H
2
S and other acid gases, must be
removed from natural gas because in the presence of water these
impurities can form acids that corrode pipelines and other
equipment. Although this paper is not focussed on the removal
of H
2
S, it should be noted that health and safety is a key driver for
removal of this highly toxic gas from sour natural gas streams.
Several of the CO
2
capture technologies described are more
selective for H
2
S than CO
2
. Furthermore, CO
2
provides no heating
value and must be removed to meet gas quality specications
before distribution to gas users. The maximum level of CO
2
permitted in natural gas transmitted to customers by pipeline is
typically less than 3% (Hubbard, 2010) although local contracts
may stipulate quality specications different to these values. The
specications of CO
2
removal from natural gas to be processed in
a cryogenic plant to produce LNG, recover liqueed petroleum gas
(LPG) or natural gas liquids (NGL) are more stringent than those
for typical gas pipelines. For example, in addition to more
extensive dehydration, the CO
2
concentration of the natural gas
should be less than 50 ppmv (Hubbard, 2010) before it enters the
cryogenic processes within the plant to avoid the formation of
dry ice.
A typical gas pipeline specication for N
2
is 3% (Kidnay and
Parrish, 2006); because N
2
is inert the main driver for its removal
from pipeline sales gas is to increase the heating value of the gas.
Nitrogen will not freeze or lead to corrosion in a cryogenic gas
plant, but a maximum concentration of 1% N
2
is often specied for
LNG, for example to avoid stratication and rollover of the liquid
product during shipping. To ensure this specication is met in
plants liquefying NG with a high N
2
content, cryogenic distillation
columns known as nitrogen rejection units (NRU) are integrated
into the process; these columns are both expensive and energy
intensive. In addition, high N
2
concentrations in natural gas
processed by an LNG plant are energetically parasitic because a
signicant amount of energy is wasted cooling the N
2
from
ambient temperatures to those of LNG (about 161 1C at atmo-
spheric pressure). Furthermore, even if the NG does not have a
high N
2
concentration, an NRU may be necessary because the
liquefaction process produces two product streams: the LNG and
an end-ash gas which is a mixture of N
2
and CH
4
vapour. This
end-ash gas can be used as fuel gas for the plant and/or blended
into sales gas destined from a conveniently located pipeline.
However, if production of end-ash gas exceeds the fuel gas
requirements and if no pipeline is available for its disposal an
NRU may be necessary to produce a stream pure enough that it
can be vented to atmosphere. Given the increasing regulation of,
and costs associated with mitigation of, greenhouse gas emissions
this relatively expensive NRU option could become increasingly
common.
This review provides an overview of conventional, developing
and novel gas processing technologies for CO
2
and N
2
removal
from natural gas, with particular attention paid to large scale LNG
production. The introduction sections provide an overview of a
typical process ow sheet and the fundamental unit operations
involved in natural gas processing. The subsequent sections of
this review describe the core concepts and industrial application
of absorption, distillation, adsorption, membrane and hydrate gas
separation technologies.
Many of the process technologies we describe for CO
2
capture
from CH
4
can also be applied to CO
2
capture from combustion ue
gases, for example in coal-red power stations. The recent
advances and future trends in technologies for capturing CO
2
at
the power plant are reviewed elsewhere by many other authors,
including Figueroa et al. (2008), MacDowell et al. (2010) and
Ebner and Ritter (2009). However, the process conditions avail-
able in the natural gas processing facility can be very different to
those conditions available for post-combustion ue gas (predo-
minantly a mixture of CO
2
, N
2
and H
2
O) treatment. The two most
signicant differences between these applications are the CO
2
partial pressure, and the level of CO
2
removal required. In the rst
case, the natural gas feed to the CO
2
removal unit is typically
available at high pressures (more than 3000 kPa) while ue gases
are typically at close to atmospheric pressure, so the driving force
for CO
2
capture from ue gas (partial pressure CO
2
typically less
than 15 kPa) is much lower than from natural gas. In the second
case the level of CO
2
removal required for natural gas production
is greater, especially for LNG production plants, than bulk separa-
tion of CO
2
from ue gas. Thus, some promising strategies for
CO
2
capture in thermal power plants such as oxy-combustion
(a modied combustion process that can produce a high CO
2
concentration ue gas, Plasynski et al., 2009) are not relevant to
CO
2
removal from natural gas.
1.1. Overview of conventional gas processing ow schemes
The major operations that can be used in natural gas proces-
sing and LNG production are shown in Fig. 1. Common process
operations include inlet gas compression, acid gas removal,
dehydration, LPG/NGL recovery and hydrocarbon dewpoint con-
trol, nitrogen rejection, outlet compression and liquefaction.
Depending on the available markets, feed gas properties, product
specications and the gas ow rate, the units identied in Fig. 1
may not all be required. For example a nitrogen rejection unit
(NRU) may not be required for the production of pipeline gas from
a feed gas containing only low N
2
concentrations, or if a high N
2
content gas can be blended with richer natural gas streams to
meet pipeline gas specications.
Table 1
Volumes of natural gas traded as pipeline and LNG (billions of cubic metres) by top
natural gas exporting and importing countries in 2010 (BP Statistical Review of
World Energy June 2011, 2011).
Top natural gas exporters
Pipeline LNG Total
1 Russian Federation 186.5 13.4 199.9
2 Norway 95.9 4.7 100.6
3 Qatar 19.2 75.8 94.9
4 Canada 92.4 0.0 92.4
5 Algeria 36.5 19.3 55.8
6 Netherlands 53.3 0.0 53.3
7 Indonesia 9.9 31.4 41.3
8 Malaysia 1.5 30.5 32.0
9 U.S. 30.3 1.6 32.0
10 Australia 0.00 25.4 25.4
Top natural gas importers
Pipeline LNG Total
1 United States 93.3 12.2 105.5
2 Japan 93.5 93.5
3 Germany 92.8 92.8
4 Italy 66.3 9.1 75.3
5 United Kingdom 35.0 18.7 53.6
6 France 35.0 13.9 48.9
7 South Korea 44.4 44.4
8 Turkey 28.8 7.9 36.7
9 Spain 8.9 27.5 36.4
10 Ukraine 33.0 33.0
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 124
In conventional gas processing H
2
S and CO
2
are removed in an
acid gas removal unit (AGRU) using aqueous amine absorption
processes. The sweetened gas leaving the amine process is saturated
with water, so typically the AGRU is located upstream of the
dehydration facilities. The nal sections of the gas liquefaction plant
(main cryogenic heat exchange (MCHE) in Fig. 1) can operate at
temperatures as low as 161 1C, and therefore, it is essential that
components that could freeze, and cause blockages in the cryogenic
equipment, at low temperatures are removed from feed to the
cryogenic plant. In addition to CO
2
, components that could freeze in
the cryogenic plant include water (typically removed to less than
0.1 ppmv), heavier parafnic hydrocarbons and aromatics such as
benzene. To protect the aluminium plate-n heat exchangers used
in the NRU the feed gas to the cryogenic plant must be free of
mercury (below 0.01 mg/Nm
3
) (Kidnay and Parrish, 2006). The gas
feed entering the (Pre-Cool section of the) liquefaction plant is
typically delivered, or compressed to, a pressure of more than
3400 kPa and at a temperature close or slightly above ambient.
The conventional method of removing N
2
from natural gas is by
cryogenic distillation; hence NRUs are usually closely integrated
within the liquefaction process. There are several variations of
cryogenic liquefaction processes including, for example, Air Products
and Chemicals (APCI) C3MR process and the ConocoPhillips Opti-
mised Cascade process. Commonly a propane refrigeration loop is
used to pre-cool the gas before it enters a main cryogenic heat
exchanger (MCHE) which might use a mixed refrigerant or a cascade
of several pure uid refrigeration cycles. If an NRU is incorporated
with the liquefaction process then it would commonly be located
downstream of the MCHE and before the nal depressurisation
stage of the LNG production process.
1.2. Overview of gas separation mechanisms
Separation processes can be designed to exploit differences in
the molecular properties or the thermodynamic and transport
properties of the components in the mixture. Molecular proper-
ties that could be exploited to achieve a separation of CO
2
, N
2
and
CH
4
include the differences in kinetic diameter, polarizability,
quadrupole and dipole moments of the molecules (Table 2).
Thermodynamic and (interphase) transport properties that could
be exploited include vapour pressure and boiling points, solubi-
lity, adsorption capacity and diffusivity. Based on these properties
of the components to be separated, the primary operations for the
separation and purication of gases apply one of the following
inherent separation mechanisms: (1) phase creation by heat trans-
fer and/or shaft work to or from the mixture, (2) absorption in a
liquid or solid sorbent, (3) adsorption on a solid, (4) permeation
through a membrane and (5) chemical conversion to another
compound (Kohl and Nielsen, 1997; Seader and Henley, 2006).
The rst four of these operations are discussed in this review;
direct chemical conversion of CO
2
from natural gas to a useful
product is beyond the current scope. An example of such direct
conversion, which is attracting signicant current research inter-
est, is the dry reforming process where CO
2
reacts with CH
4
to
produce syngas (a mixture of H
2
and CO
2
) which can then be used
for the production of liquid fuels via FischerTropsch reactions.
Recent reviews on the conversion of CO
2
include articles by
Havran et al. (2011), Song (2006) and Zangeneh et al. (2011).
Separations involving the creation of a phase include the partial
condensation or partial vaporisation of species with very different
volatility from a feed mixture; and this category can also include
Fig. 1. Conventional gas processing operations in a typical natural gas plant with a cryogenic process for liqueed natural gas production. The required operations and
process ow arrangements vary depending on the local feed gas properties, size of the plant, and product specications; and may not all be required (MCHEmain
cryogenic heat exchanger).
Table 2
Physical properties of methane, carbon dioxide and nitrogen.
Property CH
4
CO
2
N
2
Kinetic diameter (

A) (Tagliabue et al., 2009)
3.80 3.30 3.64
Normal boiling point (NBP) (K) (Lemmon et al., 2010) 111.7 77.3
Critical temperature (K) (Lemmon et al., 2010) 3.80 304.1 126.2
Critical pressure (kPa) (Lemmon et al., 2010) 4600 7380 3400
DH
vap
at NBP (kJ/mol) (Linstrom and Mallard, 2011) 8.17 26.1 5.58
Polarisability (

A
3
) (Tagliabue et al., 2009)
2.448 2.507 1.710
Quadrapole moment (D

A) (Tagliabue et al., 2009)
0.02 4.3 1.54
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 125
desublimation of CO
2
. If the volatility differences among species
to be separated are not sufciently large, as is the case for
N
2
CH
4
, to achieve the desired separation in a single partial vapor-
isation or partial condensation contact stage, then a distillation
process involving multiple vapour-liquid contact stages may be
required.
One of the most fundamental properties of an inherent separa-
tion mechanism is its selectivity with respect to the components i
and j being separated. As will be discussed, some separation
mechanisms have more than one type of selectivity; however, the
most common and applicable type is the equilibrium selectivity, a
ij
,
which is dened in terms of phase compositions:
a
ij

x
i
x
j
_ _
y
j
y
i
_ _
1
Here x
i
and x
j
are the mole fractions (or convenient concentra-
tion units) of components in one equilibrium phase, and y
i
and y
j
are the mole fraction of components in a second equilibrium
phase. This denition of equilibrium selectivity derives from its
central role in distillation theory, where it is also known as the
relative volatility (McCabe et al., 2005). However, the equilibrium
selectivity can be applied more broadly than to vapourliquid
equilibrium in distillation; it is useful in the analysis of adsorp-
tion- and absorption-based processes. The equilibrium selectivity
can also be applied, through the use of appropriate analogies and
use of concentrations of the permeate (x) and feed (y) streams, to
the analysis of membrane processes which are not strictly
equilibrium-based. It is important to note that selectivity is often
linked to the characteristic energy of the separation mechanism,
and therefore the regeneration energy needed if the mechanism is
to be used in a cyclical process.
The performance of a separation process is governed by two
factors: the inherent selectivity of the separation mechanism being
utilised and the degree to which that mechanism can be exploited
through appropriate engineering design. This gives rise to another
metric, known as the separation power, SP
ij
(Seader and Henley,
2006), which can be used to quantify the performance of an entire
separation process in terms of the product stream compositions. For
a single stage equilibriumoperation a
ij
SP
ij
; however, for a process
utilizing multiple separation stages SP
ij
values much larger than the
single stage equilibrium selectivity can be achieved.
SP
ij

C
1
i
C
1
j
_
_
_
_
C
2
j
C
2
i
_ _
2
Here C
1
i
and C
2
i
are the concentrations of component i in each of
the product streams, and C
1
j
, C
2
j
the concentrations of component j
in each of the product streams. The symbol C for concentration has
been used here to distinguish Eq. (2) as the separation power of the
process from the equilibrium selectivity of the mechanism, although
in many cases the composition of the product streams used in the
calculation of SP
ij
may be in units of mole fraction. In principle any
selective mechanism with a
ij
41 could be engineered to achieve a
given separation power if no constraints exist on capital, operational
and energy costs. In practice a separation process will be selected if it
can achieve the separation power necessary to meet the desired
product specications at a cost that is (i) lower than other processing
alternatives, and (ii) economically viable in terms of the
products value.
In Table 3, indicative values of a
ij
and SP
ij
are given for the most
common industrial processes used to separate CO
2
or N
2
from natural
gas for the purpose of evaluating the prospects of new and emerging
technologies. To generalise this measure for any of the four primary
separation operations the product streams must be dened. For CO
2
removal from natural gas, a convenient approach may be to compare
the a
CO
2
,CH
4
values of the various separation technologies by dening
the equilibrium phases to be used with Eq. (1) as the CO
2
selective
phase: that is the amine solution or physical solvent for absorption,
the high-purity liquid CO
2
product for distillation, the adsorbed phase
(x
i
or q
i
) for gassolid adsorption systems, and the permeate stream
from a CO
2
selective membrane stage. Using available data for CO
2
removal processes, such as the amine absorption data included in
Kohl and Nielsen (1997), the typical separation power of processes to
produce pipeline quality gas (2% CO
2
) from a feed mixture containing
5% CO
2
can be estimated with Eq. (2) as shown in Table 3. The
inherent equilibrium selectivity can be exploited for each separation
mechanism to achieve SP
ij
ba
ij
through the arrangement of multiple
separation stages into an engineering process system, for example:
trayed absorption and distillation columns, multiple adsorption beds
operating in adsorptiondesorption cycles, and membrane stages
with interstage recompression and recycle loops. This simplied
analysis shows clearly that the typical inherent process separation
powers for chemical absorption technologies are much larger than
the best alternatives currently available to treat large gas ows; hence
amine absorption is the most commonly applied type of AGRU
despite the energy required for regeneration and corrosive nature of
the amine solutions.
For N
2
rejection technologies the conventional cryogenic distilla-
tion technology has a separation power more than 8 times that of N
2
selective adsorption and membrane processes. However, the
Table 3
Inherent equilibrium selectivity, a
i,j
, for the separation of CO
2
from CH
4
and N
2
from CH
4
by different operations, with the typical process separation power (SP
i,j
) achieved
in example technologies implementing these separation operations. The separation powers are calculated for typical processes to (a) remove CO
2
from a feed gas
containing 5% CO
2
to produce pipeline quality gas with 2% CO
2
, and (b) to reject N
2
from a gas containing 4% N
2
to produce a stream with 1% N
2
for LNG production.
Process Separating agent Typical inherent
equilibrium selectivity
Typical process
separation power
(a) CO
2
/CH
4
separation a
CO2,CH4
SP
CO2,CH4
Amine absorption (MDEA) Liquid absorbent 860 3300
Physical solvent (chilled methanol) Liquid absorbent 318 1900
AdsorptionCO
2
selective Solid adsorbent 28.5 622
MembraneCO
2
selective Membrane 1520 2040
(b) N
2
/CH
4
separation a
N2,CH4
SP
N2,CH4
Cryogenic distillation Heat transfer 58 320
AdsorptionN
2
selective Solid adsorbent 1.32
a
840
a
AdsorptionCH
4
selective Solid adsorbent 0.250.5 1.052
MembraneN
2
selective Membrane 23 210
MembraneCH
4
selective Membrane 0.250.3 210
a
Inherent kinetic selectivities of narrow pore adsorbents are reported from 2 to 10, which could allow a N
2
selective process with much higher separation power to be
engineered.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 126
separation of N
2
from CH
4
in adsorption-based processes is
enhanced by the differences in rates of diffusion of N
2
and CH
4
that
have been reported for small pore titanosilicate ETS-4 materials
(Marathe et al., 2004a). Guild Associates Molecular Gate
TM
pressure
swing adsorption (PSA) process (Guild Associates, 2007) is a
commercial example of a N
2
rejection process that relies on the
kinetic selectivity of N
2
over CH
4
.
In practice, the achievable separation power may be much
lower than the ideal separation powers estimated in Table 3.
Furthermore, there are many other factors beside a separation
power or inherent selectivity that a process engineer designing or
selecting a gas separation process will need to consider. Other
factors that inuence the selection of process technology for CO
2
/
CH
4
and N
2
/CH
4
separations include the level of contaminants in
the feed gas, the required level of contaminant removal or
product purity, the ow rate and condition of the feed gas
(temperature, pressure, water content), and for AGRUs the need
for simultaneous or selective removal of H
2
S. Process selection is
also inuenced by the available disposal routes for the removed
contaminants, which may include reinjection of CO
2
for enhanced
oil recovery (EOR) or enhanced gas recovery (EGR), and venting of
N
2
to atmosphere. The process plant layout and available plant
space for the separation process must be considered. If the
separation process is to be installed on an offshore platform, a
oating LNG (FLNG) plant or as a retrot in an existing production
plant, then the process footprint the plan area and/or height
occupied by the process equipment may inuence the choice of
process technology. Each of these process selection criteria may
have material impacts on the feasibility, energy requirements and
costs of CO
2
and N
2
removal processes.
There are many factors that inuence the cost of a separation
process including the extent to which it has been used
successfully in the past. Consequently, a process engineer design-
ing or selecting a gas separation process is likely to be more
interested in the ratio of its separation power to its cost rather
than in the inherent selectivity upon which the process is based.
However, once a separation process is sufciently mature, its
separation power to cost ratio will generally only improve
asymptotically, unless a signicant improvement in its inherent
selectivity can be achieved. Thus, the starting point for scientists
and engineers aiming to develop a new separation technology
should be an analysis of the inherent selectivities of both the
new technology and the conventional one with which it will
compete.
2. Absorption
This section focuses primarily on CO
2
absorption processes,
but also introduces technologies for N
2
rejection by the selective
absorption of CH
4
in hydrocarbon solvents (Mehra and Gaskin,
1997) and the potential for the development of N
2
selective
solvents. Although this review is concerned primarily with CO
2
removal, the selection of the AGRU process is more often
determined by the H
2
S removal requirements. Thus, most of the
technical literature concerning acid gas treating focuses on the
absorption of H
2
S. Commonly, the capacity of a sorbent is
reported as an acid gas loading capacity which includes the
capacity for CO
2
and H
2
S. We have noted in Tables 5 and 10
which of the process technologies are suitable for the simulta-
neous or the selective absorption of H
2
S.
A large number of commercial processes are available for CO
2
absorption in chemical and physical solvents, including the
technologies listed in Table 4. Chemical absorption processes
with amine solutions are the most commonly used acid gas
removal technologies in the natural gas industry (GPSA
Engineering Data Book, 2004). The chemical absorption processes
rely on reactions of the CO
2
with the sorbent to form weakly
bonded intermediate compounds, and these reactions can be
reversed by the application of heat to release the CO
2
and
regenerate the sorbent (Olajire, 2010). Physical solvents, such as
the mixture of polyethylene glycoldimethyl ethers used in the
Selexol
s
process, selectively absorb CO
2
from the natural gas feed
according to Henrys law so that absorption capacity increases at
high pressure and low temperature.
The two major cost factors in gasliquid absorption processes
are (1) the required sorbent circulation rate, which is determined
by the amount of CO
2
that must be removed from the feed gas
and the CO
2
loading capacity of the sorbent, and (2) the energy
required to regenerate the sorbent (Kidnay and Parrish, 2006).
The acid gas loading capacity of physical solvents at low to
moderate CO
2
partial pressures is generally lower than that of
chemical absorbents. However, physical solvent processes have
lower energy requirements for regeneration because the heat of
Table 4
Examples of commercial absorption technologies for CO
2
capture and gas sweetening.
Vendor/licensor Sorbent Reference
Chemical absorption
Econamine
SM
Fluor Digylcolamine, mono-ethanolamine http://www.uor.com
ADIP-X Shell MDEAaccelerator www.shell.com
aMDEA
s
BASF MDEA www.basf.de
GAS/SPEC Ineos MDEA www.gasspec.com
UCARSOL DOW MDEA http://www.oilandgas.dow.com
KM CDR Mitsubishi Heavy Industries KS-1 hindered amine Mimura et al., (1995)
Beneld UOP Potassium carbonate UOP Overview of Gas Processing
Technologies and Applications, (2010)
Catacarb Eickmeter and Associates Potassium carbonate (organic additive) www.catacarb.com
Flexsorb HP Exxon Mobil Potassium carbonate (steric amine) www.exxonmobil.com
Physical absorption
Fluor Solvent
SM
Fluor Dry propylene carbonate www.uor.com
Selexol UOP/DOW Mixed polyethylene glycol dimethyl ethers www.uop.com
Purisol Lurgi n-Methyl-2-pyrrrolidone www.lurgi.com
Rectisol Lurgi Chilled methanol www.lurgi.com
Ifpexol IFP Chilled methanol Larue and Lebas, (1996)
Mixed-solvent processes
Sulnol-D Shell SulfolaneDIPAwater www.shell.com (Rajani, 2004)
Amisol Lurgi Methanol secondary aminewater Kohl and Nielsen, (1997)
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 127
absorption for physical solvents is much lower than the heat of
absorption for chemical solvents.
2.1. Chemical absorption processes for acid gas treating
2.1.1. Aqueous amine processes
Amines are organic compounds derived from ammonia (NH
3
)
where one or more hydrogen atoms have been substituted with
an alkyl or aromatic group. It is the (NH
2
) functional group of the
amine molecule that provides a weak base that can react with the
acid gases. The absorption of CO
2
occurs via a two-step mechan-
ism: (1) the dissolution of the gas in the aqueous solution,
followed by (2) the reaction of the weak acid solution with the
weakly basic amine. The rst physical absorption step is governed
by the partial pressure of the CO
2
in the gas feed. The reactions
involved in the second step of CO
2
absorption in aqueous amines
have been widely studied, with a large number of reference
materials on the reaction mechanisms (Bindwal et al., 2011;
Kohl and Nielsen, 1997; Penny and Ritter, 1983; Vaidya and
Kenig, 2007; Versteeg et al., 1996) and guidelines for process
operation (GPSA Engineering Data Book, 2004) available in the
literature. The fundamental reactions involved in CO
2
absorption
in amine treating are (Kohl and Nielsen, 1997):
Dissociation of water:
H
2
O"H

OH

(3)
Hydrolysis and dissociation of dissolved CO
2
:
CO
2
H
2
O"HCO
3

(4)
Protonation of the amine:
RNH
2
H

"RNH
3

(5)
Carbamate formation:
RNH
2
CO
2
"RNHCOO

(6)
The dissociation reactions are shown here to highlight that the
pH of the amine solution is an important process parameter
because the concentrations of the ionic species H

, OH
-
and HCO
3
-
in the amine solution affect the other reactions involving
the amine.
Amines can be classied according to the number of hydrogen
atoms that have been substituted, as primary (RNH
2
, where R is
a hydrocarbon chain), secondary (RNHR
0
) or ternary (R
0
NR
R
00
). For primary and secondary amines, such as monoethanola-
mine (MEA) and diethanolamine (DEA), the carbamate formation
reaction (Eq. (6)) predominates; this reaction is much faster than
the CO
2
hydrolysis reaction (Eq. (4)). The stoichiometry of the
carbamate reaction suggests that the capacity of primary and
secondary amines is limited to approximately 0.5 mol of CO
2
per
mole of amine. However, DEA-based amine processes can achieve
loadings of more than 0.5 mol of CO
2
per mole of amine through
the partial hydrolysis of carbamate (RNHCOO
-
) to bicarbonate
(HCO
3

), which regenerates some free amine (Kidnay and Parrish,


2006).
Tertiary amines such as MDEA, which do not have a free
hydrogen atom around the central nitrogen, do not react directly
with CO
2
to form carbamate. Instead, CO
2
reactions with tertiary
amines proceed via equivalent reactions to those shown in Eqs.
(4) and (5), which are much slower than the reaction in Eq. (6), to
give the overall reaction:
RR
0
R
00
NCO
2
H
2
O"RR
0
R
00
NH

HCO
3

(7)
The stoichiometry in Eq. (7) shows that theoretically tertiary
amines can achieve a loading of 1 mol of CO
2
per mole of amine,
which is double the CO
2
loading capacity of primary amines. Also
the required heat of regeneration is lower for tertiary amines. A
disadvantage of tertiary amines is that the absorption kinetics are
slower than for primary and secondary amines. For some natural
gas treating processes the slow kinetics of CO
2
absorption in
tertiary amines can be utilised to achieve selective H
2
S removal
by optimisation of the contact time in the absorber to minimise
CO
2
uptake (GPSA Engineering Data Book, 2004). Alternatively, to
enhance CO
2
the absorption kinetics of tertiary amines an
activator (usually a primary or secondary amine) may be added
to increase the rate of hydrolysis of carbamate and dissolved CO
2
(GPSA Engineering Data Book, 2004).
Since the rst application of tertiary amines in the mid-1970s,
signicant research has been directed into the further develop-
ment of novel amine solvents. To accelerate the reaction rate of
tertiary amines with CO
2
, a primary or secondary amine can
be included as an activator (GPSA Engineering Data Book, 2004).
Fig. 2. Process ow diagram of a typical amine-solvent (MDEA)-based chemical adsorption system for the separation of CO
2
and other acid gases from natural gas
(Hubbard, 2010; Kohl and Nielsen, 1997).
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 128
For example, the cyclic diamine piperazine has been studied as a
promoter to improve the CO
2
mass transfer rates of MDEA and
MEA (Bishnoi and Rochelle, 2002); a commercial example of this
technology is the aMDEA solvent from BASF. The piperazine
reacts rapidly with CO
2
in the vapour phase, which accelerates
the dissolution of CO
2
into carbonic acid, which can then react
quickly with MDEA. Sterically hindered amines are either primary
or secondary amines with large bulky alkyl or alkanol groups
attached to the nitrogen (Seagraves and Weiland, 2011), which
reduces the carbamate stability. The molecular conguration
dictates the amount of CO
2
removal: severely hindered amines,
such as ExxonMobils Flexsorb SE, have very low rates of CO
2
absorption and allows selective H
2
S removal. In contrast, moder-
ately hindered amines, such as 2-amino, 2-methyl, 1-propanol
(AMP), are characterised by high rates of CO
2
absorption and high
capacities for CO
2
. Weiland et al. (2010) evaluated the use of AMP
with MEA for CO
2
capture from ue gases and found it had several
advantages including at least a 15% reduction in the required
regeneration energy.
A typical process ow diagram for the removal of acid gas from
a sour gas feed using methyldiethanolamine (MDEA) is shown in
Fig. 2. Properties and typical operating conditions for commonly
used aqueous amine solutions are shown in Table 5. The basic
process ow for other amine absorption systems is similar to that
shown for MDEA, although some commercial process designs
often feature multiple column feeds and contactor sections. Any
liquids or solids in the sour feed gas are removed in an inlet
separator before the gas enters at the bottom of amine contactor.
Typical operating pressures for amine contactors are in the range
of 5070 bar (Kidnay and Parrish, 2006). The lean amine solution,
typically an aqueous solution containing 1065%wt amine, is fed
at the top of the column. As the amine solution falls down the
contactor and mixes with the gas, the acid gases dissolve and
react with the amine to form soluble carbonate salts. The
sweetened natural gas leaves the top of the contactor saturated
with water and so dehydration is normally required before the
gas is sold or fed to a cryogenic gas plant. Process temperatures
inside the contactor rise above ambient temperature due to the
exothermic heat of absorption and reaction, with a maximum
temperature observed near the bottom of the column.
The rich amine leaves the bottom of the contactor at a tempera-
ture of approximately 60 1C and containing 0.200.81 mol of acid
gas per mole of amine (GPSA Engineering Data Book, 2004). The
pressure of the rich amine stream is reduced to around 6 bar in a
ash tank, to separate any dissolved hydrocarbons from the rich
amine, and preheated to 80105 1C before entering the stripping or
regenerator column. In the stripping column, heat supplied by a
steam reboiler generates vapour, which removes the CO
2
from the
rich amine as the vapour travels up the column. A stream of lean
amine is removed from the bottom of the stripper, cooled to
approximately 40 1C and recycled to the amine contactor. The
vapour stream from the top of the stripping column is cooled to
condense and recover water vapour, and the acid gas may be vented,
incinerated, sent to a sulphur recovery plant (for H
2
S rich feed gas)
or compressed for reinjection into a suitable reservoir for enhanced
oil/gas recovery (Hughes et al., 2012).
The process disadvantages with conventional amine treating
processes include: (1) the large amounts of energy required for
regeneration of the amine, (2) the relatively low CO
2
loading
capacity of amines requires high solvent circulation rates and
large diameter, high-pressure absorber columns, (3) the corrosive
amine solutions induce high equipment corrosion rates, (4) degra-
dation of amines to organic acids, and (5) co-absorption of
hydrocarbon compounds such as benzene, toluene, ethylbenzene
and xylene (BTEX) which subsequently are emitted with the acid
gas stream (Collie et al., 1998; Morrow and Lunsford, 1997).
Operational issues also include solution foaming, emulsions,
excessive solution losses, heat stable salts and high-lter change
out frequency (Seagraves and Weiland, 2011). Aqueous ammonia
and hot carbonate systems are among the alternative chemical
absorption processes to amines. However, many of the disadvan-
tages of amine treating are also associated with aqueous ammo-
nia and hot carbonate processes.
Future innovations in conventional absorption column tech-
nology (e.g., tray and packing designs) could be expected to
achieve only incremental improvements in process efciencies
(MacDowell et al., 2010). However, one promising strategy to
intensify the CO
2
absorption process is the use of a hollow bre
membrane as a gasliquid contactor device (Cai et al., 2012;
Ebner and Ritter, 2009; Favre and Svendsen, 2012; Zhou et al.,
2010). In this concept, the membrane does not show any
selectivity for CO
2
over CH
4
; instead the membrane provides a
physical barrier between the gas and liquid phases, and a large
interfacial surface area for mass transfer of CO
2
. The selective
Table 5
Properties of common aqueous amine solvents for acid gas treating.
Solvent Monoethanolamine Diethanolamine Digylcolamine Methyldiethanolamine
Acronym MEA DEA DGA MDEA
Normally capable of meeting H
2
S specication
(Kidnay and Parrish, 2006)
yes yes yes yes
Removes COS, CS
2
, mercaptans (Kidnay and Parrish,
2006)
partial partial partial partial
50 ppm CO
2
for cryogenic plant feed (Kidnay and
Parrish, 2006)
no, 100 ppm possible yes, 50 ppmv in
SNEA-DEA
no, 100 ppm
possible
no, pipeline
quality only
Solvent degradation concerns (components) (Kidnay
and Parrish, 2006)
yes - COS, CO
2
, CS
2
, SO2,
SO3, mercaptans
some - COS, CO
2
,
CS
2
, HCN, mercaptans
yes - COS, CO
2
,CS
2
no
Solution concentrations, normal range wt% (GPSA
Engineering Data Book, 2004)
15-25 30-40 50-60 40-50
Acid gas pickup, mole acid gas / mole amine (GPSA
Engineering Data Book, 2004)
0.33-0.40 0.20-0.80 0.25-0.38 0.20-0.80
Rich solution acid gas loading, mol/mol amine
normal range (GPSA Engineering Data Book, 2004)
0.45-0.52 0.21-0.81 0.35-0.44 0.20-0.81
Lean solution acid gas loading, mol/mol normal
range (GPSA Engineering Data Book, 2004)
0.12 0.01 0.06 0.005-0.01
Stripper reboiler normal range, 1C (GPSA Engineering
Data Book, 2004)
107-127 110-127 121-132 110-132
Approximate integral heats of absorption of CO
2
, kJ/
mol Kohl and Nielsen, 1997
84.4 71.6 83.9 58.8
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 129
absorption of CO
2
in the liquid solvent occurs at the liquid
interface of the membrane. For the hollow bre membrane to
operate effectively as a gasliquid contact device the pores must
remain gas lled and preventing liquid penetration into the
membrane pores is one of the practical challenges hindering the
commercialisation of this technology (Favre and Svendsen, 2012).
A gasliquid membrane contactor pilot plant is reported to have
been tested in Scotland to remove CO
2
from a 5000 Nm
3
/h natural
gas feed (Mansourizadeh and Ismail, 2009). Other improvements
in the performance of CO
2
absorption processes are likely to come
through the development of new solvent materials. Some acid gas
treating process licensors are also working to develop chemical
additives to inhibit corrosion and solvent degradation, so that
amine-based solvents can be operated at higher amine concen-
trations and regenerated at higher temperatures (Normand et al.,
2012). Improved understanding of the thermodynamics and
kinetics of CO
2
-amine systems, including the development of
mass transfer rate-based modelling approaches, is also allowing
optimised design of absorption processes.
2.1.2. Hot carbonate (alkali salt) systems
Technologies using hot solutions of potassium carbonate
(K
2
CO
3
) or sodium carbonate (NaCO
3
) have been employed since
the 1950s to remove CO
2
from high pressure gas streams (Kohl
and Nielsen, 1997). The overall reactions for CO
2
with potassium
carbonate can be represented by (GPSA Engineering Data Book,
2004):
CO
2(g)
K
2
CO
3
H
2
O
(l)
"2KHCO
3(s)
(8)
The basic potassium carbonate process was developed by the
US Bureau of Mines and commercialised as the Beneld Process in
1954, and now licensed by UOP with over 700 units constructed
(UOP Overview of Gas Processing Technologies and Applications,
2010). Other commercial potassium carbonate technologies com-
peting with the Beneld Process include the Catacarb Process
(Eickmeter and Associates), which is mainly used in the ammonia
industry, and the Flexsorb HP Process (Exxon Research and
Engineering).
The process ow diagram for a potassium carbonate absorp-
tion system shares many features with the general amine process
ow diagram shown in Fig. 2. In a typical hot carbonate process
design the absorber and stripping columns operate in a tempera-
ture range of 100116 1C (GPSA Engineering Data Book, 2004). If
H
2
S removal is required or if low CO
2
concentrations are required
in the product gas, then alternative designs with a two-stage
contactor and a lean-solution pumped to the middle of the
absorber may be used. For gas treatment requiring CO
2
removal
to low levels for cryogenic gas processing, or GTL plant feed, the
UOP process design can be modied to a Hi-Pure design that
combines the potassium carbonate process and an amine process
(UOP Overview of Gas Processing Technologies and Applications,
2010; Miller et al., 1999). Amines such as DEA and MEA are also
used as activators to increase the rate of absorption of CO
2
in the
potassium carbonate solution (Kohl and Nielsen, 1997). The
Catacarb Process is characterised by the use of a proprietary
organic additives to improve mass transfer rates (Kohl and
Nielsen, 1997). The distinguishing feature of the Flexsorb HP
Process is the use of a sterically hindered amine as the activator
(Kohl and Nielsen, 1997) which is claimed to improve CO
2
loading
capacity and mass transfer rates.
2.1.3. Aqueous ammonia solvents
Similar to the acid gas absorption processes using aqueous
amine solutions, processes based on the reaction of CO
2
with
ammonia (NH
3
) in solution have been developed for the capture
of CO
2
from natural gas, coal seam gas, and post combustion ue
gases (Darde et al., 2010; Gonzalez-Garza et al., 2009). There are
two variants of the aqueous ammonia process (AAP) reported:
(i) chilled AAP designs operating with absorber temperatures in
the range 020 1C and (ii) processes operating with absorber at
ambient temperatures (2540 1C). Both variants are based on the
same reactions of CO
2
and ammonia (NH
3
) described by Bai and
Yeh (1997), but at low temperatures the chilled AAP allows the
precipitation of ammonium bicarbonate shown in Eq. (9):
CO
2(g)
NH
3(l)
H
2
O
(l)
"NH
4
HCO
3(s)
(9)
A further advantage of the chilled AAP design is that absorber
operation at low temperatures reduces ammonia slip into the
sweetened gas.
The process ow scheme of an AAP plant is very similar to the
ow scheme for amine absorption cycles (Fig. 2) with an absorp-
tion column and a solvent regeneration system. The absorption
column in the AAP operates at low pressures, usually close to
ambient, and the CO
2
-rich slurry leaving the bottom of the
absorber column must be pumped to a high-pressure, high
temperature regeneration column (Gal, 2006). In the regeneration
column the ammonium bicarbonate solid can be decomposed to
NH
3
and CO
2
at temperatures greater than about 50 1C (Darde
et al., 2010; Olajire, 2010), although temperatures of 100150 1C
are preferred in some designs (Gal, 2006). Typical AAP solvent
concentrations are in the range 1330% wt NH
3
(Kim et al., 2008;
Olajire, 2010).
Although the aqueous ammonia process has potentially lower
energy requirements than amine absorption processes (one study
on AAP for postcombustion CO
2
capture suggests 21003100 kJ/
kg CO
2
compared to 3700 kJ/kg for CO
2
capture by MEA, Darde
et al., 2010), the energy savings are not sufciently large to offset
the additional costs associated with complexity of the ammonia
process (Kohl and Nielsen, 1997) and the need to recompress the
sweetened gas in AAPs. Also, the removal efciency of chilled
AAPs is only 90% (Gal, 2006), hence ammonia processes may not
be capable of achieving very low CO
2
concentrations in product
gas required for cryogenic gas processing.
2.2. Physical solvent and hybrid solvent processes
Physical solvent processes may be competitive with amine
absorption when the feed gas is available at high pressure
(generally greater than about 20 bar) or when the acid gas partial
pressure is 10 bar or greater (Nichols et al., 2009). For onshore
natural gas processing facilities all the commercial physical
solvents listed in Table 4 could be used for bulk removal of CO
2
.
Due to their large plant footprints, physical solvent technologies
are generally not suitable for AGRUs on offshore facilities (Nichols
et al., 2009). To treat feed gas with very high CO
2
concentrations,
the leading physical absorption technologies include the Selexol
s
and Rectisol
s
processes (Burr and Lyddon, 2008).
The regeneration of physical solvents can be achieved by
reducing the pressure of the rich solvent stream in a series of
multi-stage ash vessels, as shown in Fig. 3, or by stripping the
absorbed gas species in a regeneration column. Importantly, the
heat inputs required for regeneration of a physical solvent are
generally much lower than the heat required for the regeneration
of amine or potassium carbonate sorbents. A potential short-
coming of low pressure regeneration cycles is the cost of recom-
pressing the acid gas if it is to be further processed for CO
2
sequestration, EOR or sulphur recovery.
The Selexol
s
process, based on a mixture of polyethylene
glycol-dimethyl ethers, is able to remove CO
2
simultaneously
with H
2
S and water (GPSA Engineering Data Book, 2004). In fact,
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 130
H
2
S has a greater solubility in most organic solvents than CO
2
, a
property that can be used to design H
2
S selective processes (Kohl
and Nielsen, 1997). As many physical solvents also absorb water,
in contrast to the aqueous amine, AAP and potassium carbonate
technologies which saturate the sweet gas with water, the
required capacity of any dehydration units downstream of the
AGRU may be smaller if a physical absorption process is used.
To produce a sweet gas containing less than 50 ppmv CO
2
for
feed to a LNG plant, the Rectisol
s
process using a methanol
solvent operating at temperatures as low as 35 to 75 1C has
been applied successfully (GPSA Engineering Data Book, 2004). A
reported example of a chilled methanol plant is at Riley Ridge,
Wyoming, where 200 MMscfd of ultra-rich CO
2
gas (70%) is
treated (CO
2
Extraction & Sequestration Project Riley Ridge, WY,
2011).
The main weakness of physical solvent technologies relative to
amine AGRUs remains the issue of relatively low acid gas
adsorption capacities of the commercially available physical
solvents. Consequently physical solvent circulation rates are high;
thus large diameter absorption columns and solvent circulation
equipment are required in physical solvent processes. For onshore
gas processing facilities, the capital costs associated with the high
solvent circulation rates may be at least partially offset by the
lower costs of the carbon steel materials required when using
non-corrosive physical solvents compared to the more expensive
materials required to handle the highly corrosive aqueous amine
solutions (GPSA Engineering Data Book, 2004).
Several mixed-solvent (also known as hybrid solvent) gas
treating processes combine the effects of physical and chemical
absorption processes in a single operation. The most well-known
mixed-solvent processes for CO
2
absorption are the Sulnol-D
s
process licensed by Shell Global Solutions (Rajani, 2004) and the
Amisol process licensed by Lurgi (Kohl and Nielsen, 1997). The
Sulnol-D
s
process based on a mixture of Sulfolane (tetrahy-
drothiophene dioxide), DIPA (diisopropanolamine) and water is
capable of deep removal of CO
2
to less than 50 ppm from feed gas
containing very high concentrations of CO
2
(Kohl and Nielsen,
1997). Another variation of the Shell Global Solutions technology
is the Sulnol-M
s
process, which is also based on Sulfolane but
uses MDEA instead of DIPA, used mainly for selective removal of
H
2
S. The process ow scheme of the Sulnol-D
s
process is
essentially the same as that for an amine absorption process
(Fig. 2) with the addition of a ash tank to remove the bulk of the
acid gas from the rich solvent upstream of the stripper column.
The Amisol process is based on a mixture of methanol, water and
either diethylamine (DETA) or diisopropylamine (DIPAM). This
process has most commonly been used for purication of synth-
esis gas derived from coal, peat, or heavy oils (Kohl and Nielsen,
1997). The advantages of the hybrid solvent technologies over
conventional amine absorption technologies include low energy
consumption for regeneration of the solvent, high acid gas loading
capacities, low foaming tendency, and reduced corrosion. Hybrid
solvent processes are usually only suited for treatment of natural
gas with an acid gas partial pressure of more than 100 kPa. The
main drawback of the mixed-solvent processes is that hydrocar-
bon losses (to the solvent) are slightly higher than the typical
losses in conventional amine processes.
2.3. Ionic liquids and switchable solvents
Among the materials investigated as new solvents for CO
2
absorption processes, ionic liquids (ILs) are one of the solvents
that may in the future offer an alternative to amines and the low
capacity physical solvents. Ionic liquids are commonly dened as
organic salts with melting temperatures of less than 373 K. They
have a range of properties that may make them useful replace-
ments for volatile organic solvents such as extremely low vapour
pressure, nonammability, and in many cases low toxicity (Zhao,
2006). Furthermore, ILs are often considered to be designer
solvents because of the many different possible combinations of
cations and anions which can be used to tune their chemical and
physical properties (Brennecke and Gurkan, 2010). These proper-
ties of ILs have generated great scientic interest in their devel-
opment and investigation into their use in chemical engineering
applications, including gas separations. Reviews by Zhao (2006),
Plechkova and Seddon (2008) and Werner et al. (2010) describe
the present industrial processes which use ionic liquids and
discuss many other possible industrial applications of ionic
liquids. The application of ILs to CO
2
capture and natural gas
sweetening has been discussed by many authors, including Bara
et al. (2010b), Karadas et al. (2010), Brennecke and Gurkan (2010)
and MacDowell et al. (2010).
The development of ILs for CO
2
capture and natural gas sweet-
ening has been driven by the desire to develop a solvent with a CO
2
capacity comparable to that of amine-based solvents but with
greatly reduced energy requirements for regeneration. Initially,
much of the research focussed on the potential of imidazolium-
based ILs as alternative physical solvents. Early studies focussed on
Fig. 3. Process ow diagram of a typical physical solvent process for absorption of CO
2
and other acid gases from natural gas.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 131
the most convenient ILs to synthesise, such as 1-n-alkyl-3-methy-
limidazolium [C
n
mim] cations paired with various anions including
bis(triuoromethylsulfonyl)amide [Tf
2
N] (Bara et al., 2010b; Hughes
et al., 2011). However, the alkyl chains (in the imidazolium cation)
are not the ideal functional group for separating CO
2
from CH
4
or N
2
and, thus, these alkyl chains were substituted with various groups
containing either ethylene glycol or nitrile units to form
[Rmim][Tf2N] (Bara et al., 2010b). Karadas et al. (2010) summarised
the results of using anions other than [Tf
2
N] on the solubility of CO
2
in various ILs, with anions containing uorinated derivatives show-
ing a modest increase.
Despite these efforts to optimise the molecule, the solubilities
achieved with ILs on a volumetric basis (dissolved moles of CO
2
per
liquid volume) have remained comparable with most common
organic solvents (Bara et al., 2010b; Karadas et al., 2010). Much
research has thus focussed on the development of task specic ionic
liquids (TSIL) which incorporate an amine functional group into the
IL, enabling it to serve as a reactive, chemical solvent for CO
2
. Bates
et al. (2002) reported an imidazolium-based TSIL containing an amine
functional group attached to one of the alkyl chains. The stoichio-
metry of the absorption reaction achieved was 0.5 mol of CO
2
per
mole of TSIL; regeneration of the IL solvent was also achieved by
heating the carbamate product to 80100 1C under vacuum. Gurkan
et al. (2010) and Zhang et al. (2009) used TSILs with amino-acid
functional groups that improve the stoichiometry to nearly 1 mol of
CO
2
per mole of IL, which is important if the IL solvent is to be
competitive with the volumetric capacities of aqueous amine solu-
tions (Brennecke and Gurkan, 2010). However, TSILs can be difcult
to synthesise and have large viscosities at ambient temperature,
which increase further upon complexation with CO
2
(Brennecke and
Gurkan, 2010; Karadas et al., 2010). Brennecke and co-workers
suggest that the viscosity increase of TSILs upon complexation with
CO
2
can be eliminated through the use of aprotic heterocyclic anions
and have led a provisional patent (Brennecke and Gurkan, 2010).
Nevertheless, the synthesis and viscosity challenges associated with
TSILs currently limit their commercial viability.
Currently, the most viable method of applying ILs to CO
2
capture
or natural gas sweetening is the use of ILamine mixtures, in which
MEA or DEA is dissolved in an [Rmim][Tf2N] solvent; such solutions
can have up to 116 times the CO
2
solubility on a volumetric basis
than the IL alone (Bara et al., 2010a). Camper et al. (2008) found that
MEA-IL and DEA-IL solutions could rapidly and reversibly capture
1 mol of CO
2
per mole of amine and thereby reduce feed gas CO
2
concentrations to the ppm level, even at CO
2
partial pressures below
0.133 kPa. Such ILamine solvents offer signicantly reduced
energy requirements relative to conventional aqueous amine sol-
vents: Bara et al. (2010a) compared a IL-amine process with ash
regeneration (similar to the process ow scheme shown in Fig. 3),
against a conventional amine-based gas sweetening plant (similar to
that shown in Fig. 2). They considered the sweetening of 100
MMSCFD to a sales gas specication of 2% CO
2
for an NG feed
containing either 15% or 5% CO
2
and concluded that over a 20 yr
plant life cycle, the amineIL process had combined CAPEX and
OPEX savings of about 25%. This was due primarily to the removal of
a regeneration column, the higher amine loadings and lower solvent
circulation rates of the amineIL solvent, and the reduced duties
required to cool, heat and regenerate the amineIL solvent. The
difference in the caloric properties of water and the IL, and in
particular the duty reduction associated with not vaporizing any
solvent during regeneration, is probably the most signicant advan-
tage of the amineIL solvent approach to gas sweetening. A pilot
scale unit is under construction for planned NG sweetening eld
tests in 20112012 (Bara et al., 2010a).
Brennecke and Gurkan (2010) point out that all ILs are
hygroscopic (that is, a material that can adsorb or absorb water
molecules), even those that are usually designated hydrophobic
because they are insoluble in water. This fact represents a serious
challenge for the application of ILs to CCS but it also has
implications for NG sweetening. If the NG contains some water,
then the absorption of that water by the IL could decrease the
solvents capacity, and will also degrade the reduction in regen-
eration duty associated with an amineIL solution. Thus in
contrast with conventional gas processing practice, a sweetening
process utilising an amineIL solvent should probably be situ-
ated downstream of the dehydration process.
Jessop and co-workers reported the development of switch-
able solvents (Jessop et al., 2005; Phan et al., 2008) where a basic,
non-polar liquid mixture converts into a polar IL upon the
addition of CO
2
. Several groups are researching the optimisation
of these switchable solvents for improved CO
2
capture or NG
sweetening. For example, Heldebrant et al. (2011) reported the
development of 2nd generation switchable solvents that could
capture of nearly 1.3 mol of CO
2
per mole of solvent. In these
cases, the viscosity of resulting IL is appreciable and its reduction
is the focus of ongoing research.
2.4. Absorption processes for N
2
rejection
Nitrogen rejection using absorption-based technologies is not
a common practice in the natural gas industry. There are several
commercial N
2
rejection processes that operate by the physical
absorption of CH
4
in a hydrocarbon oil which have been built to
process gas feed rates of 230 MMscfd (AET-Technology, 2007;
TGPE, 2009). The costs of large solvent rates and CH
4
recompres-
sion are prohibitively high for the use of CH
4
selective absorption
NRUs in large scale LNG production plants; thus there is a
growing need for the development of a N
2
selective absorbent.
An example of a CH
4
selective NRU process is that designed
and constructed by Advanced Extraction Technologies, Inc. (AET,
Houston USA). The ow scheme of the AET process is similar to
the physical solvent process shown in Fig. 3. In a typical CH
4
selective absorption process the feed gas is cooled before entering
the absorption column where CH
4
is dissolved into a lean oil
solvent (TGPE, 2009). The CH
4
is then recovered from the rich
solvent through a series of ash vessels to produce sales gas
containing less than 4% N
2
. The N
2
from the feed gas ows
through the top of the column and is used for precooling the unit
feed, and is then vented. The AET process operates at approxi-
mately 32 1C to optimise the CH
4
absorption in the solvent.
However, the capacity of most solvents for CH
4
is relatively low
and large solvent circulation rates are required to achieve eco-
nomic recovery of the CH
4
. Furthermore, because the CH
4
is
recovered from the solvent by ashing at low pressure the gas
must be recompressed for pipeline or cryogenic gas plant feed.
Although most solvents have a greater solubility for CH
4
than
N
2
, some organo-metallic complexes (OMCs) have been reported
to preferentially bind N
2
from natural gas mixtures. Reversible
transition metal complexes forming with N
2
were rst reported in
the scientic literature by Allen and Senoff (1965) who described
the reversible reaction of hydrazine (a N
2
source) with ruthenium
chloride. Several scholarly articles describing the preparation and
N
2
absorption capacity of these types of transition metals forming
complexes were also published in the 1970s (Allen et al., 1973;
Chatt et al., 1978; Sellmann, 1974). However, there have been few
studies reported on the application of these N
2
binding complexes
to natural gas processing. The most extensive applied studies of
N
2
selective solvents are reported by Stanford Research Institute
International (SRI), on contract for the US Department of Energy
(Alvarado et al., 1996), and Bend Research Inc. (Friesen et al.,
2000; Friesen et al., 1993). Both the SRI process and the Bend
process exploit the reversible chemical complexing abilities of
multi-dentate transition metal complexes as shown in Fig. 4.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 132
The absorption of N
2
in OMCs is achieved through using metal
ions with six coordination sites, the four equatorial sites complexed
with bi, tri, or tetra dentate ligands; with an electron withdrawing
ion on one axial site leaves the other axial site ready to complex
with a gaseous N
2
molecule in an end-on conguration. The strength
and therefore reversibility of the complex formation is dictated
through the use of different ligands and transition metals. The most
promising OMC reported by SRI was a (bis)tricyclohexylphosphine
molybdenum tricarbonyl, in a toluene solution, which formed a
yellow precipitate when bound with a N
2
molecule (Bomberger
et al., 1999). This phosphine complex absorbed up to 0.12 mmol of
N
2
/mL of solution at 1979 kPa (Alvarado et al., 1996) in equilibrium
measurements. However, during absorptionregeneration cycles
measured on a 0.02 MMscfd bench-scale apparatus the SRI research-
ers encountered issues with the degradation of the phosphine
complex and regeneration of the complex to release the N
2
. The
problems with the regeneration of the phosphine complex included
difculties in controlling the size of the N
2
-bound precipitates,
which would bypass the regeneration system if the size was less
than that of the phosphine solids (412 mm) (Bomberger et al.,
1999). Although SRIs economic analysis suggests that this OMC
process could become cost competitive with cryogenic NRUs, the
challenges to reduce the cost of OMC synthesis, improve the stability
of the phosphine complex in the presence of water and oxygen, and
to overcome the problems encountered with solids handling in the
regeneration process are all signicant.
The patents held by Bend Research Inc. (Friesen et al., 2000,
1993) describe OMCs based on several transition metalligand
combinations including complexes with iron, as well as the perfor-
mance of these complexes in N
2
absorptionregeneration cycles. The
patents report N
2
uptakes of 0.5 mol N
2
/mol OMC at a N
2
partial
pressure of approximately 1130 kPa and a selectivity for N
2
over CH
4
close to 6. These complexes are reported to exhibit a high degree of
stability, showing no decrease in capacity through repeated use over
100 days, and exposure to an atmosphere of 3% CO
2
and 100 ppm
O
2
. Although the publically available information on the Bend
process indicates the concept of N
2
absorption in these OMCs may
be sound, no commercial process has been developed for this
technology. The main barriers to the use of OMCs in large scale
gas processing operations remain the high cost of synthesis of the
OMCs and improving the chemical resistance of the complexes to
common gas contaminants such as water and H
2
S. Furthermore, the
Bend Research Inc. iron-based OMCs may present safety and
materials handling issues because these are pyrophoric compounds,
which can ignite spontaneously on contact with air.
3. Condensation, desublimation and distillation
3.1. N
2
rejection by cryogenic distillation
The normal boiling point (NBP) of CH
4
is 161.5 1C and at a
typical pressure of 3150 kPa in the intermediate stages of the
cryogenic LNG plant the boiling point (BP) of CH
4
is 94.7 1C
(Lemmon et al., 2010), which provides a sufcient difference in
relative volatility with N
2
(NBP 195.8 1C, BP of 148.7 1C at
3150 kPa) for separation of CH
4
N
2
mixtures by cryogenic
distillation. Relevant to such separations are vapourliquid equi-
librium curves, such as the one shown in Fig. 5 for N
2
CH
4
at
conditions accessible in a cryogenic LNG plant. The equilibrium
curve in Fig. 5 represents a typical relative volatility a
N
2
,CH
4
value
of 58. To illustrate the separation of N
2
and CH
4
in the cryogenic
distillation column, we have applied the McCabeThiele method
(McCabe and Thiele, 1925) to estimate the number of ideal
equilibrium stages required to produce a liquid CH
4
bottoms
product of 95% from a feed containing 50% N
2
and 50% CH
4
. These
feed and production compositions were selected here to allow
illustration of the separation process; in a real process to produce
a CH
4
containing less than 1% N
2
and with a high rate of CH
4
recovery, the number of actual equilibrium stages required in the
Fig. 4. General chemical scheme for the reversible binding of nitrogen with an organo-metallic complex adsorption (Miller et al., 2002). Rgeneric organic functionality.
Fig. 5. Illustration of the binary CH
4
N
2
vapourliquid equilibrium relationship
and the construction of a McCabeThiele diagram to calculate the number of ideal
equilibrium stages for separation by distillation. The vapourliquid equilibrium
data shown here represents a relative volatility, a
N2CH4
, of 7 (calculated using
REFPROP (Lemmon et al., 2010) for an operating pressure of 2757 kPa. In this
hypothetical example, the feed contains 50% N
2
with product specications of 5%
N
2
in the CH
4
-rich bottom product and 5% CH
4
in the N
2
-rich overheads product.
These compositions were selected as an example which could be shown clearly on
this gure and do not represent a typical set of operating conditions. To meet more
realistic processing objectives, such as a higher purity CH
4
-rich bottom product
and 98% recovery of the methane from the feed gas, a much larger number of
equilibrium stages would be required.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 133
distillation column would be higher than the eight ideal trays
shown in the McCabeThiele construction on Fig. 5. (The corre-
sponding graph would be more difcult to read.) In practice,
modern cryogenic NRUs (with SP
N
2
,CH
4
320) can produce very
high purity CH
4
at high recovery rates, which also reduces the CH
4
content in the N
2
overheads vapour to less than 3%. This is
achieved through the use of columns with large numbers of
stages and the design of systems with multiple columns.
Currently, cryogenic distillation is the only N
2
rejection method
that has been demonstrated at gas ows above 25 MMscfd to
achieve very high methane recovery (typically above 98%) and high
purity N
2
(approximately 1% CH
4
). The selection, and optimised
design, of a cryogenic NRU is principally determined by the
concentration of N
2
in the feed gas (Wilkinson and Johnson, 2010).
The feed gas pressure, feed gas ow rate, concentration of con-
taminants and product specications (sales gas or LNG) also
inuence process selections. The optimum design of cryogenic
NRUs, like for most cryogenic processes, is an exercise in balancing
the energy efciency and process ow sheet integration to reduce
the power consumption required for compression of the CH
4
refrigerant loops, which are used to provide the reboiler and
condenser duties (Finn, 2007; Wilkinson and Johnson, 2010). The
compression requirements of this distillation-based separation are
the largest contributor to capital and operating costs of the NRU
process. Cryogenic NRU processes have been constructed by most of
the major process designers such as Linde, Costain, Praxair, Con-
ocoPhillips and APCI. The main variants of cryogenic NRU designs
are (1) the single-column heat-pumped process, (2) the double-
column process (Agrawal et al., 2003) and (3) three or two column
designs featuring a prefractionation column (MacKenzie et al., 2002;
Wilkinson and Johnson, 2010).
A typical single-column heat-pumped NRU process is illu-
strated in Fig. 6. Upstream of the cryogenic NRU, contaminants
that could freeze at cryogenic temperature such as water, CO
2
and
heavy hydrocarbons have been removed from the gas. The feed
gas to the NRU is cooled, throttled and fed to an intermediate
stage of the distillation column operating at pressures from 1300
2800 kPa (Agrawal et al., 2003). Rejected N
2
vapour (typically
o1% CH
4
) is drawn from the column overheads and the CH
4
-rich
liquid product is drawn from the bottom of the column. The
bottoms product then can be reheated against the NRU feed gas.
A closed-loop CH
4
heat-pump cycle driven by an external com-
pressor provides the reboiler and condenser duties, with the
closed-loop CH
4
condensed at a high pressure in the reboiler
and revaporised at low pressure in the condenser.
As the N
2
content in the feed gas increases, the CH
4
in the upper
stages of the column becomes more difcult to condense. The
operating exibility of a single-column NRU process is limited by
(1) the critical pressures of nitrogenmethane mixtures, which
limits the maximum pressure of the distillation column to approxi-
mately 2800 kPa, and (2) the minimum practical temperature of CH
4
after the throttling valve of the heat-pump cycle (MacKenzie et al.,
2002). These limitations mean that the single-column NRU process
is, generally, used for feed gases containing less than 20% N
2
.
A double-column NRU can provide additional process exibility
compared to the single-column process to allow the separation of
gases containing higher N
2
concentrations or gases in which the feed
gas quality varies. In the double-column N
2
rejection process the
NRU feed gas is cooled, throttled and fed to a high pressure (HP)
column operating typically at 10002500 kPa (Agrawal et al., 2003).
Having had some of the N
2
removed, the crude natural gas liquid
stream from the bottoms of the high pressure column is sub-cooled,
throttled and fed to the low pressure (LP) column (operating at
approximately 150 kPa). In practice both the HP and LP columns are
usually integrated into a single tower to improve process heat
integration and minimise heat transfer to the atmosphere. The
N
2
-rich vapour from the HP column is condensed to provide reux
for both the high pressure column and the low pressure column. The
low pressure column produces a high purity N
2
stream (o1% CH
4
);
which is used to cool N
2
reux fed to the LP column in the LP
condenser, sub-cool the bottoms of the HP column and to pre-cool
the feed gas. The CH
4
-rich bottom product of the LP column is
pumped through the crude sub-cooler, which provides the LP
columns reboiler duty by condensing the overheads of the HP
column, vaporised and reheated against the feed gas. Alternatively,
for LNG production the CH
4
-rich bottoms product of the LP column
may remain liquid.
The three-column NRU process consists of the double-column
process described and a prefractionation column. The prefractio-
nation column recovers some of the hydrocarbons at a higher
temperature than the double-column system and increases the N
2
concentration of the feed gas. Importantly, the prefractionation
column reduces the volume of N
2
-rich gas that must be processed
at low temperatures in the main NRU column(s) and this reduc-
tion in gas volume can signicantly reduce the power require-
ments of the NRU (MacKenzie et al., 2002; Wilkinson and
Johnson, 2010). A further variant of an NRU with a prefractionator
described by MacKenzie et al. (2002) is a two column process
with a high pressure prefractionation column, an intermediate
liquidvapour separator and a low pressure column.
3.2. Low-temperature CO
2
removal processes
The separation of CO
2
from natural gas by low-temperature
processes (operating at temperatures below 0 1C) can be
Fig. 6. A process ow schematic of a typical single-column N
2
rejection unit
(schematic adapted from GPSA Engineering Data Book, 2004; Agrawal et al., 2003;
MacKenzie et al., 2002).
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 134
categorised as: (1) gasliquid phase separations operating at
temperatures above the CO
2
triple point temperature of
56.6 1C (Lemmon et al., 2010) and (2) gassolid phase separa-
tions where desublimation of CO
2
occurs at temperatures below
the triple point. Although the term cryogenic is often used by
the vendors and in the literature to describe these types of CO
2
capture technologies, most of the processes operate at tempera-
tures above the scientic denition of cryogenic as 153 1C
(Agrawal et al., 2003; Radebaugh, 2007). To overcome the
problems associated with the formation of CO
2
solids during
cryogenic distillation two technological approaches have been
pursued: (1) extractive distillation by the addition of a heavier
hydrocarbon to alter the solubility of components in the column
(Ryan/Holmes process) and (2) controlled freezing and re-melting
of the solids (Controlled Freeze Zone
TM
and CryoCell
s
processes).
Other low-temperature CO
2
removal technologies under devel-
opment include systems in which mechanical methods are used
to separate the CO
2
rich phase from the natural gas. For example,
Willems et al. (2010) report the C3sep (condensed contaminant
centrifugal separation) process in which condensed CO
2
droplets
are separated from the natural gas using rotational separators,
and Clodic et al. (2005) describe the ALSTOM process which
features a multi-stage thermal swing process that freezes then
melts CO
2
on mechanical ns. We focus our discussions in this
review on the commercialised Ryan/Homes process and the pilot-
plant demonstrated Controlled Freeze Zone
TM
(along with the
similar CryoCell
s
process).
The extractive distillation approach to solving the problem of
CO
2
freezing in CH
4
CO
2
distillation is most well known through
the Ryan/Holmes process described in the 1982 US Patent
4,318,723 (Holmes and Ryan, 1982). The Ryan/Holmes process
is representative of several similar technologies patented by
various other inventors. The addition of a heavier hydrocarbon
stream (typically a C
2
C
5
alkane) to the condenser of the distilla-
tion column shifts the operation away from conditions that favour
solids formation, because the solubility of CO
2
in the liquid phase
can be increased, the overheads temperature can be raised, and
the column can be operated at a higher pressure since the
mixtures critical pressure increases. A typical four column
Ryan/Holmes process conguration incorporates a de-ethaniser
column, a CO
2
recovery column, a demethaniser column, and a
column for recovery of the hydrocarbon additive. Further details
on Ryan/Holmes congurations and operating issues for separa-
tions of methaneCO
2
, ethaneCO
2
, and CO
2
H
2
S are discussed in
the GPA Engineering Data Book.
The Controlled Freeze Zone
TM
(CFZ
TM
) process was rst patented
by ExxonMobil in 1985 (Valencia and Denton, 1983) and tested in a
Texas pilot plant during 19861987 (Nichols et al., 2009). More
recently a commercial demonstration project designed to treat a
feed gas of 14 MMscfd has been constructed in LaBarge, Wyoming
(Controlled Freeze Zone
TM
increasing the supply of clean burning
natural gas, 2010). The separation tower of the CFZ
TM
process is split
into three sections with an upper rectication section and a lower
stripping section (both conventional distillation sections) separated
by the CFZ
TM
section, as shown in Fig. 7 (Fieler et al., 2008). In the
CFZ
TM
section, the liquid falling from the rectication section is
contacted with a cold methane stream (90 to 85 1C), which
causes the CO
2
to freeze out of the methane mixture. The CO
2
solids
(62 to 45 1C) drop to a liquid layer on a melt tray in the lower
stripping section; the solids melt before falling as liquid through the
downcomers of the melt tray. The standard CFZ
TM
process can
produce pipeline quality gas, and when implemented with a
modied rectication section is claimed to be capable of producing
a sweet gas of less than 50 ppm CO
2
(Nichols et al., 2009).
Cool Energys CryoCell
s
was developed by researchers at
Curtin University in Western Australia with industrial partners
Woodside Petroleum and Shell Global Solutions (Hart and
Gnanendran, 2009). Like the CFZ
TM
process, the CryoCell
s
process
operates by the controlled freezing and subsequent remelting of
CO
2
. The basic thermodynamic path for the CryoCell
s
operation
involves cooling a dry, feed gas (at 56006600 kPa) to just above
the CO
2
freezing point (for example to 60 1C) to condense some
or all of the vapour, followed by an isenthalpic ash to further
cool the mixture to obtain solids, liquid CO
2
and a CH
4
-rich
vapour. Pilot plant trials of the CryoCell
s
process demonstrated
the production of pipeline quality gas from 2 MMscfd of feed
gases containing 3.560% CO
2
(Hart and Gnanendran, 2009). The
cryogenic separation of CO
2
has potential as a highly selective
process to treat CO
2
-rich natural gas, although research efforts to
overcome the operational issues associated with control of the
CO
2
freezing and solids handling were underway in 2009.
The key advantages of phase creation processes for CO
2
separation from CH
4
over amine-based absorption systems for
separation of CO
2
from CH
4
include the recovery of a high purity,
liquid CO
2
product at a reasonable pressure, which facilitates the
subsequent transport or injection for use in EOR; the avoidance of
highly corrosive aqueous amine solvents, and possibly, reduced
process footprint and reduced hydrocarbon inventories, which
may be important considerations for offshore or oating produc-
tion facilities (Kelley et al., 2011).
4. Adsorption
The separation and purication of gas mixtures by the selec-
tive adsorption of components from the gas mixture onto porous
solid adsorbents is an established process technology used in
the production of hydrogen, the separation of O
2
and N
2
from air,
and the capture of odorous pollutants from various industrial
processes. In the natural gas industry adsorption-based separa-
tions are used to remove water, sulphur, mercury and heavy
Fig. 7. General schematic of the Controlled Freeze Zone
TM
process with spray
nozzles in the CFZ section (schematic adapted from Fieler et al., 2008).
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 135
hydrocarbons (for dew point control) from the natural gas
(Tagliabue et al., 2009). Adsorption-based processes for the
separation of mixtures of CH
4
, N
2
and CO
2
are also used for
post-combustion CO
2
capture (Ebner and Ritter, 2009), the pur-
ication of coal mine methane (Richter et al., 1985; Tonkovich,
2004; US Environmental Protection Agency, 2008; US EPA, 1997)
and coal mine ventilation air (VAM) (Warmuzinski, 2008), and for
the purication of biogas (Alonso-Vicario et al., 2010; Esteves
et al., 2008). Central to the development and implementation of
adsorption-based processes are the various selectivity mechan-
isms that give rise to the separation of components within the gas
mixture.
4.1. Adsorbent selectivity
The preferential adsorption of components from a gas mixture
can be achieved by one, or a combination, of the following
mechanisms: (1) differences in the adsorbatesurface interactions
and/or adsorbate packing interactions when the system reaches
equilibrium (thermodynamic equilibrium mechanism), (2) differences
in the size and/or shape of gas molecules leading to exclusion of
molecules with a critical diameter too large to enter the adsorbent
pores (steric mechanism) and (3) differences in the diffusion rates of
molecules through the adsorbent pores (kinetic mechanism) (Li et al.,
2009; Ruthven, 2011). The kinetic mechanism can include the
quantum sieving effect of different diffusion rates observed for
some light molecules in narrow micropores (Xiao-Zhong et al.,
2009). Most industrial adsorption processes such as NG dehydration
using silica desiccants or molecular sieves (Kohl and Nielsen, 1997)
rely on the thermodynamic equilibrium effect. The separation of N
2
from CH
4
in the Molecular Gate
TM
PSA process using ETS-4 and the
separation of O
2
and N
2
from air using carbon molecular sieves and
small pore zeolites (Kerry, 2007) are industrial examples based on
the kinetic mechanism. True steric or size exclusion-based processes
are unlikely to be viable technologies for CO
2
/CH
4
or N
2
/CH
4
separations because the differences in the critical diameters of
CO
2
, N
2
and CH
4
are not sufciently large for total exclusion of
one component fromthe adsorbent pores. Based on these adsorptive
gas separation mechanisms there are two types of adsorbate
selectivitythe equilibrium selectivity achieved in the limit of long
time periods and the kinetic selectivity (or time dependent selectiv-
ity). The equilibrium selectivity a
ij
of the adsorption mechanism
dened in Eq. (1) can be written for an equilibrium selective
adsorption process as:
a
ij

q
i
=y
i
q
j
=y
j

K
i
K
j
as y
i
,y
j
-0 10
where q
i
and q
j
are the equilibrium adsorption capacities deter-
mined from pure gas component isotherms, and y
i
and y
j
are the
mole fractions of the components in the gas mixture. In many cases,
including at low partial pressures of the component species, the
separation factor can be estimated as the ratio of the Henrys
constants (K
i
/K
j
) (Tagliabue et al., 2009). This denition of a
ij
serves
as a useful tool to screen potential adsorbents for CO
2
and N
2
removal from natural gas. However, to adequately design an
adsorption-based separation process the selectivity of the adsorbent
for components from a real gas mixture must be conrmed and the
working capacity of the adsorbent needs to be evaluated (Ackley
et al., 2003). The working capacity is the difference between the
amounts of a component adsorbed and desorbed at the conditions of
the adsorption and desorption steps, and this capacity is inuenced
strongly by temperature and pressure.
The equilibrium capacity for a gas species is inuenced by the
strength of the gassolid interaction and the number of available
adsorption sites. The strength of the gassolid interaction is
determined by the characteristics of the adsorbents surface
chemistry and pore structure; and by the adsorbates properties
including molecule size, polarizability and quadrupole moments.
Typical heats of adsorption for CH
4
and N
2
on commercial
adsorbents are in the range of 1522 kJ/mol (Cavenati et al.,
2004; Watson et al., 2009; Xu et al., 2008). Carbon dioxide
exhibits a large quadrupole moment, thus adsorbents with polar
surfaces that have a high electric-eld gradient, such as zeolites,
have a stronger interaction with CO
2
than with the non-polar CH
4
and N
2
molecules (Li et al., 2009). For example, Cavenati et al.
(2004) report the isosteric heat of adsorption of CO
2
on zeolite
13X is 37.2 kJ/mol, Xu et al. (2008) report 49.9 kJ/mol on Na b-
zeolite, and Watson et al. (2012) report 44.9 kJ/mol on a natural
chabazite. Fig. 8 illustrates the differences in equilibrium adsorp-
tion capacity at 298 K for CO
2
, N
2
and CH
4
on zeolite 13X. For
selective adsorption of a non-polar molecule like CH
4
, an adsor-
bent with a high surface area (such as microporous carbons)
possessing a large number of adsorption sites is likely to be a good
candidate.
For gas separations based on differences in sorption rates a
kinetic selectivity factor b
ij
which incorporates the effects of each
components sorption mass transfer coefcient k
i
can be dened
as follows:
b
ij
a
ij

k
i
k
j

-
y
j
y
i
b
ij

q
i
q
j
_ _
k
i
k
j

11
As discussed recently by Ruthven (2011) the kinetic selectivity
depends on both the diffusivity ratio (assuming k
i
pD
c,i
, where
D
c,i
is the diffusivity coefcient of component i in adsorbent
pores) and the equilibrium selectivity, which if being inferred
from pure uid measurements can be estimated from the second
equality in Eq. (11). Materials that exhibit a kinetic selectivity for
CO
2
or N
2
from CH
4
include carbon molecular sieves (Bae and Lee,
2005; Cavenati et al., 2005), CuMOF (Bao et al., 2011a) and
small-pore zeolites such as clinoptilolite (Ackley and Yang, 1991;
Herna ndez-Huesca et al., 1999).
4.2. Adsorption-based separation processes
Similar to the solvent absorption processes, adsorption-based
processes for gas separation require both adsorption and
Fig. 8. Equilibrium adsorption capacity of CO
2
, CH
4
and N
2
at 298 K on zeolite 13X.
Figure constructed from data reported in Cavenati et al. (2004).
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 136
regeneration stages. Adsorbent regeneration, or desorption, can
be achieved by utilizing the differences in adsorption capacities at
different temperatures (thermal-swing adsorption, TSA) and at
different pressures (pressure-swing adsorption, PSA), as illu-
strated in Fig. 9. Continuous TSA and PSA processes operate with
multiple beds containing a stationary adsorbent and use a mani-
fold of valves to switch gas ow to the beds corresponding to
adsorption and desorption cycles. Less commonly used technol-
ogies for continuous adsorption processes are uidised and
moving bed operations (Seader and Henley, 2006), and xed-
bed electrothermal-swing adsorption (ESA) (An et al., 2011;
Grande and Rodrigues, 2008).
In the TSA method, the adsorbent is regenerated by desorption
at a higher temperature than that used during the adsorption
phase of the cycle. In the natural gas industry TSA processes with
silica gel or zeolite molecular sieve lled adsorbent beds have
been used widely for gas dehydration. The temperature of the bed
can be increased by purging the bed with a hot, inert and non-
adsorbing gas, or less commonly by heat transfer from heating
coils located within the bed. After desorption the bed temperature
is reduced with a cool purge gas, and the adsorption cycle starts
again. In gas dehydration TSA units, heating and cooling the bed
can take several hours (or even days). The adsorbents used for
dehydration have a large capacity and high selectivity for water
compared to other natural gas components. However, if the
selectivity of the adsorbent for the contaminant is not so strong
the bed will become saturated quickly. To treat a large volumetric
gas ow with a long cycle TSA process will then require a bed
with a large total adsorption capacity, and thus a large amount of
adsorbent. In the PSA method, adsorption occurs at elevated
pressure (typically in the range 4002000 kPa) and desorption
occurs at near-ambient pressure (Seader and Henley, 2006). Using
the PSA method adsorbent beds can be depressurised and re-
pressurised rapidly, allowing cycle times of several minutes or
even several seconds to be utilised. Accordingly, the amount of
adsorbent required for PSA processes can be much smaller than
for an equivalent TSA processes.
Examples of commercially available PSA-based systems for
CO
2
and N
2
separation from natural gas are listed in Table 6. Most
adsorption-based CO
2
capture technologies are limited to
processing natural gas feeds containing no more than 2% CO
2
because the quantity of adsorbent required to capture greater
volumes of CO
2
is large. Recent advances in small-footprint PSA
systems for CO
2
removal from NG on offshore platforms include a
1 MMscfd Molecular Gate system operated by the Tidelands Oil
Company (Wills and Mitariten, 2009) and the 2.5 MMscfd Xebec
rapid cycle PSA system at a Veneco Inc eld (Toreja et al., 2011),
both in California.
Commercial PSA N
2
rejection processes shown in Table 6
include UOPs Nitrex
TM
process (UOP, 2010), the Nitrotec process
(Richter et al., 1985), and a micro-scale NRU designed by
American Energies Pipeline (AEP) (2009) (Nhattacharya et al.,
2009). The Nitrex
TM
process developed by UOP uses the proprie-
tary Polybed
TM
PSA platform and, in the 1990s, small scale NG
treating units based on this process were commissioned in Texas
(Tagliabue et al., 2009) to process about 2.3 MMscfd. At least two
Nitrotec plants were built in Texas during the 1990s to treat
natural gas ows of 15 MMscfd each (Tagliabue et al., 2009). The
micro-scale AEP system for processing 0.0750.5 MMscfd of low
BTU gas in Kansas includes two adsorbent towers (4
0
8
0
up to
6
0
20
0
) lled with activated carbon granules (an off-the-shelf
commercial adsorbent) (American Energies Pipeline, 2009;
Nhattacharya et al., 2009). These processes operate by the
selective adsorption of CH
4
on activated carbons or zeolites such
as 13X or 5A, and each of these systems requires a compressor to
return the CH
4
rich stream to pipeline delivery pressure. The one
commercial N
2
rejection technology that operates by the selective
adsorption of N
2
is Guild Associates Molecular Gate
TM
PSA
process (Guild Associates, 2007), which relies on the kinetic
selectivity of N
2
over CH
4
on proprietary adsorbents based on
synthetic titanosilicate ETS-4 and contracted titanosilicate 1 (CTS-1)
(Butwell et al., 2002a; Kuznicki, 1990; Kuznicki et al., 1999; Kuznicki
et al., 2000).
Adsorption-based processes have potential for energy and
capital investment savings over the conventional CO
2
amine
scrubbing and cryogenic distillation NRU technologies, but to
date adsorption-based technologies have been limited to proces-
sing natural gas feed rates of only about 15 MMscfd. However,
other industries operate much larger scale modern commercial
cyclic adsorption processes: for example, 200 MMscfd PSA plants
Fig. 9. Schematic representation of a two-bed cyclic adsorption process with pressure-swing adsorption (PSA) and thermal-swing adsorption (TSA) cycles shown on
isotherms for CO
2
adsorption on a carbon molecular sieve Takeda MSC 3K-171 (Watson et al., 2009).
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 137
are used in the production of high purity H
2
from steam methane
reformers (Ritter and Ebner, 2007). The two key conceptual
strategies that may help develop adsorption-based processes for
CO
2
and N
2
removal from natural gas at LNG scales or for gas
elds with very high contaminant concentrations are (i) modi-
cations to PSA and TSA process congurations and (ii) improved
performance, cost and reliability of adsorbents.
4.2.1. Adsorption cycle process renements
Modern PSA and TSA processes have expanded on the two-
bed, four-step conguration of early Skarstrom (1960) cycles
(pressurisation, feed gas adsorption, depressurisation, purge) to
include additional process steps to maximise productivity and
energy savings. Advanced PSA designs include the use of three or
more beds, multi-layered adsorbent beds to remove different
impurities (Cavenati et al., 2006), pre-treatment or guard beds to
remove strongly adsorbed species, and tanks for storing inter-
mediate process streams between cycle steps to use as purge and
repressurisation gases. Novel PSA cycle designs may feature more
than 10 process steps including multiple pressurisation steps, co-
current and counter-current depressurisation steps, and several
purge steps; the optimum design of such complex cycles requires
new mathematical approaches (Mehrotra et al., 2011) and experi-
mental validation in pilot-scale PSA plants (Zhang and Webley,
2008).
The bed adsorption and desorption cycle times have a strong
inuence on the product purity and ow. The short cycle duration
advantages of PSA processes over TSA processes can be further
enhanced by new hardware technologies that allow rapid cycle
changes. The two areas of development that can allow operation
of rapid cycle PSA systems are structured adsorbents and new
valve technologies. Operation with rapid swings in pressure can
result in the uidisation of small adsorbent pellet or granular
particles. This in turn causes mechanical attrition of the adsorbent
reducing performance, which reduces performance, increases bed
pressure drops, and the solid nes produced can damage valves
and downstream equipment. Structured adsorbents featuring
nanoparticles or nanocrystals of the microporous adsorbent
material (pore widths o2 nm), such as zeolites or active carbon,
on a support material like a honeycomb monolith or cloth sheet
containing macroporous channels (widths450 nm) (Alcaniz-
Monge et al., 2010; Ribeiro et al., 2008; Thiruvenkatachari et al.,
2009; Vargas et al., 2011) can overcome uidisation and pressure
drop problems in rapid swing operations. Furthermore, structured
adsorbent packings can facilitate improved mass and heat trans-
fer within the solid bed, reduce bed pressure drop and allow for a
greater volumetric density of the adsorbent than can be achieved
with pellets (Rezaei and Webley, 2010). Fast gas transport
through the adsorption bed is critical when the duration of the
adsorption process step duration approaches the time scale for
mass transfer within a single adsorbent pellet (Todd, 2003). In
practice, however, valve operation and maintenance issues can
limit the number and frequency of cycle steps.
The development of new valve technologies such as the rotary-
valves used in the Xebec M-3100 Rotary-valve fast cycle PSA
system (Toreja et al., 2011) can help to overcome limitations with
the rate at which conventional gas manifold valves can switch.
Xebec reports that their PSA system can operate at up to 50 cycles
per minute. The increased cycle rates allow more compact PSA
modules, and the rotary-valve design reduces the size of the
switching valve manifolds as well as providing operational ex-
ibility through control of the rotation speed, so this technology is
particularly useful for offshore CO
2
capture processes. Xebecs
rotary-valve, rapid cycle PSA technology has been applied in
biogas purication, H
2
purication and more recently in an
offshore plant for removal of CO
2
from natural gas. In 2010 a
Xebec M-3100 Rotary-cycle rapid cycle PSA system was commis-
sioned at the Platform Gail in the Santa Barbara Channel to
process a 2.5 MMscfd gas feed containing 13% CO
2
to a sales gas
containing 1.42% CO
2
(Toreja et al., 2011). The platform receives
gas containing 4000 ppm H
2
S, which is removed in a SulFerox
process upstream of the PSA unit. The PSA unit consists of six
adsorbent beds lled with an undisclosed non-silica, metal-based
adsorbent. The size of this rapid cycle PSA unit (18 ft(L)
8 ft(W) 9 ft(H)) is considerably smaller than the alternative of
a conventional amine-based CO
2
scrubber.
The process conditions available within a LNG plant provide
opportunities for new PSA processes to be designed to operate in
a completely different region of thermodynamic space than most
of the current commercial PSA process used in industry, which
typically operate at temperatures close to, or slightly above,
ambient. A key process advantage in an LNG plant is the avail-
ability of the feed gas at relatively high pressures and the
availability of low temperature streams which could be used for
PSA refrigeration. In particular, a refrigerant loop within the LNG
plant could conveniently be used for cooling of the PSA feed gas
approximately 243 K, and there is reason to expect that PSA
techniques for N
2
and CO
2
separation from CH
4
will be more
effective at low temperatures than at conventional temperatures
(Habgood, 1958). At low temperatures adsorption capacities are
greater and the differences between rates of adsorption of gas
components in narrow adsorbent pores could potentially be
greater at a cooler temperature (Herna ndez-Huesca et al., 1999).
Furthermore, at low temperatures for some materials the differ-
ences in rates of sorption for gas components may be exaggerated
Table 6
Survey of pressure-swing adsorption processes for CO
2
and N
2
separations from CH
4
in operation or in patent claims.
Vendor/licensor Adsorbent More adsorbed
component
Less adsorbed
component
Flow rate
(MMscfd)
Refs.
M3100 Xebec Rotary-valve Xebec Metal-based CO
2
CH
4
o2.5 Toreja et al. (2011)
CO
2
Sponge IACX Energy Activated carbon CO
2
CH
4
0.25 CO
2
Sponge (2011)
Molecular Gate Guild Associates Titanosilicates CO
2
CH
4
0.550 Guild Associates (2007)
UOP MOLSIV UOP Zeolite CO
2
CH
4
UOP (2010)
Polybed PSA/Nitrex UOP Zeolite CH
4
N
2
2.3 UOP (2010)
Micro-scale N
2
rejection unit American Energies Corporation Activated carbon CH
4
N
2
0.0750.5 (American Energies Pipeline, 2009;
Nhattacharya et al. (2009)
Nitrogen rejection unit TGPE Activated carbon CH
4
N
2
0.000315 TGPE (2009)
Nitrogen Sponge IACX Energy Activated carbon CH
4
N
2
0.35 Reinhold (2010)
Nitrotec CMS Energy Activated carbon CH
4
N
2
15 Richter et al. (1985) and
Tagliabue et al. (2009)
Molecular Gate Guild Associates/BASF Titanosilicates N
2
CH
4
0.510 Guild Associates (2007),
Mitariten (2009)
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 138
by changes in the crystalline structure of the adsorbent which
could cause changes in pore aperture size as described by Breck
(1964, 1974). Optimum design of PSA processes to operate at
novel (PSA) conditions for an LNG application require the collec-
tion and validation of experimental data for adsorbent perfor-
mance at low temperature, high pressures and measured with gas
mixtures representative of the industrial application (Jensen et al.,
2012; Watson et al., 2012).
4.3. Adsorbent state-of-the art
The desirable characteristics for an adsorbent to separate CO
2
or N
2
from natural gas include (i) a high selectivity and good
adsorption capacity for the target gas component, (ii) fast adsorp-
tion and desorption kinetics, (iii) good physical and chemical
stability through adsorption/desorption cycles and (iv) be regen-
erable by modest pressure or temperature swings to minimize
operational energy costs. The selected adsorbent must also show
robust performance in the presence of moisture and other con-
taminants that may be in the natural gas feed to the adsorption
treating unit. The cost of the adsorbent and the bed packing
density (which inuences the adsorbent bed size) are further
considerations.
4.3.1. Adsorbents for CO
2
capture
Commercial adsorbents being used to remove CO
2
from
industrial gas streams include zeolites, activated carbons and
titanosilicate molecular sieves (Table 6 lists examples of adsor-
bents used in commercial PSA systems.) Table 7 provides a survey
of the equilibrium CO
2
and CH
4
capacities, and CO
2
/CH
4
selectivities, of various commercial and novel adsorbents.
Novel-structured materials developed over the past 20 yr that
show potential for higher CO
2
capacities include adsorbents based
on metal-organic frameworks (MOFs), zeolitic imidazolate frame-
works (ZIFs), surface functionalised silicas and porous carbons
(Hao et al., 2011). Metal oxides (CaO, activated alumina, lithium
zirconate), layered hydroxides and hydrotalcites have been stu-
died extensively as sorbents for CO
2
capture from high tempera-
ture ue gases (300700 1C) (Abanades et al., 2004; Liu et al.,
2009); these high temperature sorbents are not discussed further
in this review because these temperatures are beyond the
expected range of operating conditions for CO
2
removal units in
a natural gas or LNG production plant.
Metal-organic frameworks are crystalline materials with high
internal surface areas (for example, MOF-5 has an apparent
surface area of 2900 m
2
g
1
, Eddaoudi et al., 2000) and large
pore volumes, capable of achieving high CO
2
adsorption capa-
cities. At pressures from 1040 bar CO
2
capacities of more than
10 mmol g
1
have been reported for MOF-5 (Saha et al., 2010)
and MOFs with coordinatively unsaturated metal sites (Dietzel
et al., 2009). The porous structures of MOFs can be systematically
tuned through considered selection of the metal-ions and organic
linkers that form their coordination networks (Li et al., 2009) to
improve the selectivity for target gas molecules. Carbon dioxide
selective narrow pore MOFs have been prepared using, for
example, Li (Bae et al., 2011), Mg (Mallick et al., 2010) or Cu
(Bao et al., 2011a) ions. A copperMOF prepared by Bao and co-
workers (2011a), with narrow pore apertures of approximately
0.350.35 nm, showed a CO
2
/CH
4
kinetic selectivity of 9.7 (or a
value of 25 if calculated using Henry law constants), which is one
Table 7
Equilibrium capacity, equilibrium selectivity and kinetic selectivity of adsorbents for CO
2
and CH
4
at 100 and 1000 kPa and ambient temperature.
Adsorbent name Type CO
2
capacity
(mol kg
1
)
CH
4
capacity
(mol kg
1
)
CO
2
/CH
4
equilibrium
selectivity
CO
2
/CH
4
kinetic
selectivity
T (K) P (kPa) Refs.
Mg-MOF-74 MOF 8.50 1.00 8.50 3.8 298 100 Bao et al. (2011b)
5A (Sinopec) Zeolite 4.55 0.88 5.19 3.6 298 100 Saha et al. (2010)
13X (Sinopec) Zeolite 3.30 0.37 8.92 4.5 298 100 Bao et al. (2011b)
Natural chabazite Zeolite 3.30 0.89 3.71 2.5 303 100 Jensen et al. (2011)
Na-Beta Zeolite 2.70 0.60 4.50 303 100 Xu et al. (2008)
TRI-PE-MCM-41 Amine-
silica
2.50 0.10 25.00 298 100 Belmabkhout et al. (2009)
PCB, Calgon Corp. AC 2.41 0.73 3.29 296 100 Ritter and Yang (1987)
Norit RB1 extra AC 2.20 1.10 2.00 298 100 Dreisbach et al. (1999)
MSC-3K-161 CMS 2.15 1.01 2.13 298 100 Watson et al. (2009)
PET-DC-0 AC/CMS 2.08 0.02 100.00 100.0 298 100 Cansado et al. (2010)
Sutcliffe Speakman carbon AC 1.83 0.71 2.58 298 100 Esteves et al. (2008)
CORK-DC-0 AC/CMS 1.81 0.38 4.76 8.9 298 100 Cansado et al. (2010)
H-Beta Zeolite 1.70 0.40 4.25 303 100 Xu et al. (2008)
Beta Zeolite 1.67 0.32 5.18 308 100 Huang et al. (2009)
MOF-177 MOF 1.59 0.63 2.55 3.2 298 100 Saha et al. (2010)
MSC-3K-162 CMS 1.30 0.70 1.86 303 100 Bae and Lee (2005)
ZIF-100 ZIF 1.05 0.30 3.50 298 100 Wang et al. (2008)
MOF-5 MOF 0.91 0.13 7.27 5.8 298 100 Saha et al. (2010)
Cu-MOF MOF 0.65 0.35 1.86 9.7 298 100 Bao et al. (2011a)
Maxsorb AC 13 6 2.17 298 Himeno et al. (2005)
Sutcliffe Speakman carbon AC 8.01 3.69 2.17 298 1000 Esteves et al. (2008)
Norit RB1 extra AC 7.60 4.00 1.90 298 1000 Dreisbach et al. (1999)
PCB, Calgon Corp. AC 7.20 3.90 1.85 296 1000 Ritter and Yang (1987)
13X Zeolite 6.40 2.90 2.21 298 1000 Cavenati et al. (2004)
1C
0
-Li MOF 5.30 2.40 2.21 298 1000 Bae et al. (2011)
2L
0
-Li MOF 4.60 2.50 1.84 298 1000 Bae et al. (2011)
MSC-3K-161 CMS 4.03 2.73 1.48 298 1000 Watson et al. (2009)
Natural chabazite Zeolite 3.95 1.69 2.34 303 1000 Watson et al. (2011)
MSC-3K-162 CMS 3.20 2.10 1.52 48.2 303 1000 Bae and Lee (2005)
H-Mordenite Zeolite 3.10 1.70 1.82 293 1000 Delgado et al. (2006)
Beta Zeolite 3.05 1.44 2.12 308 1000 Huang et al. (2009)
Na-Mordenite Zeolite 3.00 1.70 1.76 293 1000 Delgado et al. (2006)
1M
0
-Li MOF 2.30 0.80 2.88 298 1000 Bae et al. (2011)
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 139
of the highest selectivities reported for CO
2
/CH
4
separations.
Although a variety of MOFs show good CO
2
adsorption capacity
and excellent selectivity, at present their industrial application
may be limited by their powder form, high cost and small
available quantities, and poor chemical and thermal stability.
In addition to the widely available synthetic zeolites such as
13X and 4A, a new class of zeolitic materials known as zeolitic
imidazolate frameworks (ZIFs) are emerging as potential high
capacity CO
2
adsorbents. The framework of ZIFs is formed from
tetrahedral metal ions (for example, Zn, Co) and imidazolate
organic bridges. This structure allows ZIFs to be prepared with
tailored pore structures, similar to the preparation of MOFs, but
the chemistry also provides good chemical and thermal stability
like traditional zeolites. Wang and co-workers report a CO
2
/CH
4
selectivity for ZIF-100 of 5.9 (with 1.05 mmol/g CO
2
adsorbed at
800 Torr, 298 K) (Wang et al., 2008).
Mesoporous silica materials such as MCM-41 (Beck et al.,
1992) have been widely used for industrial gas separation, as
well as for catalyst supports for gas phase reactions. The channels
in these mesoporous silica materials such as MCM-41 and SBA-15
(Zhao et al., 1998) facilitate rapid gas diffusion, but the afnity of
CO
2
with silica surfaces is not as strong compared to the inter-
action of CO
2
with cationic sites of zeolites and MOFs. The CO
2
capacity of mesoporous silicas can be enhanced by grafting amine
groups to the hydroxyl sites on the silica surface (Belmabkhout
et al., 2009; Gray et al., 2005; Xu et al., 2002). Likewise, the CO
2
capture capacity of carbonaceous adsorbents can also be
enhanced by (i) grafting amine functional groups onto activated
carbon surfaces, or (ii) nitrogen enrichment via ammoxiation of
activated carbons (Plaza et al., 2010) or the utilisation of nitrogen
rich carbon precursors including melamine resins (Drage et al.,
2007), polypyrrole (Sevilla et al., 2011) or agricultural by-pro-
ducts such as soybean waste (Thote et al., 2010).
4.3.2. Adsorbents for N
2
rejection
A survey of reported CH
4
and N
2
adsorption capacities of a
range of commercial and novel adsorbents at 100 kPa and
ambient temperatures is shown in Table 8. This pure gas equili-
brium adsorption data is widely available for many materials, but
it should be noted that such experimental data allows an initial
ranking of adsorbents only; kinetic data for both pure uids and
gas mixtures are required for a full assessment of the adsorbents
potential for a real pressure swing adsorption process. Most
adsorbent materials show equilibrium selectivity for CH
4
over
N
2
. The methane selective activated carbons used in the PSA-
based N
2
rejection technologies described by American Energies
Corporation, TGPE, IACX Energy, and CMS Energy (in the Berg-
werksverband patent, Richter et al., 1985) are standard commer-
cial activated carbon grades, such as steam activated carbons.
Typical commercial activated carbons that are used for CH
4
selective N
2
rejection systems have (BET) surface areas of 800
1200 m
2
g
1
.
As well as the adsorbents specic surface area, the size of the
pores is also important. Several experimental and theoretical
studies identied that CH
4
adsorption capacity correlates with
micropore volume (Alcaniz-Monge et al., 2009; Contreras et al.,
2009; Kluson et al., 2000): for example, Kluson et al. (2000)
Table 8
Capacities and selectivities of adsorbents for CH
4
and N
2
at 100 kPa and 1000 kPa.
Adsorbent name Type CH
4
capacity
(mol kg
1
)
N
2
capacity
(mol kg
1
)
CH
4
/N
2
equilibrium
selectivity
CH
4
/N
2
kinetic
selectivity
T (K) P (kPa) Refs.
Sr-ETS-4, activated 543 K Titanosilicate 1.30 2.00 0.65 0.17 273 100 Marathe et al. (2004b)
G2X7/12 (Takeda) AC 1.18 0.38 3.11 298 100 Olajossy et al. (2003)
Norit RB1 extra AC 1.10 0.39 2.77 298 100 Dreisbach et al. (1999)
AX21 (MAST) AC 1.09 0.31 3.52 293 100 Scaife et al. (2000)
PET-DC-9-CVD AC/CMS 1.06 0.25 4.24 4.63 298 100 Cansado et al. (2010)
WS42 (Chemviron) AC 1.02 0.39 2.62 303 100 Belmabkhout et al. (2004)
MSC 3K-161 CMS 1.01 0.36 4.25 0.37 298 100 Watson et al. (2009), Cavenati et al. (2005)
BPL (Calgon) AC 1.01 0.34 2.97 298 100 Belmabkhout et al. (2004)
Columbia Grade L AC 0.94 0.34 2.81 303 100 Valenzuela and Myers, (1989)
Natural chabazite Zeolite 0.89 0.45 1.96 1.86 303 100 Jensen et al. (2011)
Maxsorb AC 0.88 0.32 2.76 300 100 Sheikh et al. (1996)
5A (Sinopec) Zeolite 0.84 0.56 1.50 298 100 Saha et al. (2010)
Mg-clinoptilolite Zeolite 0.80 0.45 1.78 0.10 295 100 (Jayaraman et al., 2004)
Na-Mordenite Zeolite 0.75 0.44 1.70 303 100 Delgado et al. (2006)
Sutcliffe Speakman Carbon AC 0.71 0.27 3.41 298 100 Esteves et al. (2008)
MOF-177 MOF 0.67 0.17 3.94 298 100 Saha et al. (2010)
NaX (Linde) Zeolite 0.65 0.41 1.59 305 100 Dunne et al. (1996)
Na-Beta Zeolite 0.63 0.41 1.54 303 100 Xu et al. (2008)
13X (CECA) Zeolite 0.59 0.28 2.11 298 100 Cavenati et al. (2004)
5A (WR Grace & Davison) Zeolite 0.58 0.20 2.90 303 100 Nam et al. (2005)
Sr-ETS-4, activated 373 K Titanosilicate 0.22 0.20 1.09 323 100 Cavenati et al. (2009)
Ca-ETS-4, activated 543 K Titanosilicate 0.20 0.22 0.83 0.14 323 100 Cavenati et al. (2009)
MOF-5 MOF 0.13 0.11 1.18 298 100 Saha et al. (2010)
Sr-ETS-4, activated 543 K Titanosilicate 0.10 1.46 0.07 298 100 Kuznicki et al. (2001)
Sr-ETS-4, activated 588 K Titanosilicate 0.02 0.24 0.08 0.05 295 100 Jayaraman et al. (2004)
PET-DC-0 AC/CMS 0.02 0.08 0.26 0.02 298 100 Cansado et al. (2010)
Norit RB1 extra AC 4.00 2.02 1.98 298 1000 Dreisbach et al. (1999)
Sutcliffe Speakman Carbon AC 3.69 1.82 2.03 298 1000 Esteves et al. (2008)
Columbia Grade L AC 3.47 1.78 1.95 303 1000 Valenzuela and Myers (1989)
13X (CECA) Zeolite 3.06 1.83 1.67 298 1000 Cavenati et al. (2004)
MSC 3K-161 CMS 2.73 1.33 2.05 0.18 298 1000 Watson et al. (2009), Cavenati et al. (2005)
MSC-3K-162 CMS 2.10 1.40 1.50 0.17 303 1000 Bae and Lee (2005)
Na-Mordenite Zeolite 1.70 1.50 1.13 303 1000 Delgado et al. (2006)
Natural chabazite Zeolite 1.69 1.46 1.16 303 1000 Watson et al. (2011)
Sr-ETS-4, activated 543 K Titanosilicate 0.90 3.80 0.24 298 1000 Kuznicki et al. (2001)
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 140
reported that the optimum pore width for CH
4
selective adsorp-
tion over N
2
on carbon adsorbents is 871

A (based on Density
Functional Theory and Ideal Adsorbed Solution Theory models).
The optimum pore size of 871

A allows the adsorption of a single
layer of methane within the pore. From the available data on CH
4
and N
2
adsorption capacities surveyed in Table 8, the best
materials exhibit (pure gas) CH
4
/N
2
selectivities above 4, which
is consistent with the predictions of Kluson et al. (2000) for a
typical commercial activated carbon (AX21).
Although most commercial adsorbents show equilibrium
selectivity for CH
4
over N
2
, there are adsorbents that show kinetic
selectivity for N
2
. There are even adsorbents reported such as
certain Sr-exchanged ETS-4 that exhibit equilibrium selectivity
for N
2
. Restricted diffusion of CH
4
, relative to the rate of N
2
diffusion, has been observed in materials with pores narrower
than 6

A including small pore zeolites, carbon molecular sieves
and titanosilicates. The difference in the kinetic diameters of N
2
(3.64

A) and CH
4
(3.8

A) molecules is small, so a separation
process based on the differences in the diffusion rates of N
2
and
CH
4
in an adsorbents pores requires a well-controlled, narrow
pore size distribution. The most industrially signicant examples
for N
2
rejection from natural gas are the synthetic titanosilicates
ETS-4 and CTS-1 (Butwell et al., 2002b; Kuznicki et al., 2001;
Kuznicki et al., 1999, 2000) used in the Molecular Gate
TM
PSA
technologies. The structure of ETS-4 is an interconnected
octahedraltetrahedral (with TiO
6
, SiO
4
and TiO
5
units) frame-
work with narrow 8 member ring pore openings (Kuznicki et al.,
2001). The size of the 8 member ring openings can be controlled
by exchanging the framework sodium cations for other atoms. For
example barium- (Kuznicki et al., 1999) and calcium-exchanged
ETS-4s (Cavenati et al., 2009) were reported to have enhanced
N
2
/CH
4
selectivities over the Na-ETS-4 form. Likewise, controlled
dehydration of the ETS-4 at temperatures from 373573 K can
affect the selectivity of the adsorbent as shown by the differences
in selectivity for the ETS-4 materials listed in Table 8.
Small pore zeolites have crystalline structures that may also
favour kinetic selectivity for N
2
from natural gas and the most
widely studied small pore zeolite specically for this purpose is
clinoptilolite (Ackley and Yang, 1991; Faghihian and Pirouzi,
2009; Herna ndez-Huesca et al., 1999; Jayaraman et al., 2004).
Although volumetric experiments to measure equilibrium
adsorption capacities of pure gases CH
4
and N
2
show that MSC
3K-161 has a selectivity for CH
4
at very long adsorption times
(Watson et al., 2009), the kinetic selectivity of MSC 3K-161 for N
2
over CH
4
has been observed in volumetric adsorption experi-
ments by Bae and Lee (2005) and Cavenati et al. (2005).
5. Membranes
Membrane technologies potentially offer signicant advan-
tages over traditional gas separation operations like gasliquid
absorption and cryogenic distillation. The features that make
membrane technologies highly attractive for process separation
units include the ability to separate chemical species without a
phase change, low thermal energy requirements, simple process
ow schemes with few pieces of rotating equipment, compact
plant footprints and convenient start up and shutdown proce-
dures. These features of membranes systems are potentially
attractive for remote, unmanned and footprint conscious sites.
The most successful industrial applications of membrane tech-
nologies have been conned to liquid separations such as the
purication and desalination of water. Membrane separation
technologies have been applied since the 1980s in the natural
gas industry to remove CO
2
, N
2
, H
2
S and NGLs (examples are
given in Table 9); however, membrane separation technologies
still account for less than 5% of the market for new natural gas
processing equipment installed (Baker and Lokhandwala, 2008;
Laverty and OHair, 1990). At least one plant uses a membrane
operation for helium extraction from natural gas (Laverty and
OHair, 1990).
Of the membrane technologies deployed in the natural gas
industry, processes to capture CO
2
have been the most widely
used and, currently, CO
2
capture is the only natural gas separation
process for which membrane processes are competitive with the
conventional technology (in this case amine absorption) (Baker
and Lokhandwala, 2008). This section explores the recent devel-
opments in membrane technologies that seek to overcome the
key materials performance and process challenges contributing to
the limited deployment of membrane technologies for natural gas
processing.
5.1. Membrane classications and permeation theory
Membranes for gas separation typically fall into three cate-
gories dened by their materials of manufacture: polymeric,
inorganic and mixed matrix membranes. Membrane technologies
may also be categorised by the mechanism of gas transport
through the membrane, for example, sorptiondiffusion, solution
diffusion and molecular sieving (Koros and Mahajan, 2000). The
commercial membrane technologies employed in the natural gas
industry are predominantly nonporous polymeric membranes
that separate components by the solution-diffusion mechanism
(Membrane Technology & Research, 2010). Transport of gas
molecules through the membrane by the solution-diffusion
mechanism occurs rst by absorption of the gas molecule into
the membrane and then by diffusion, or permeation, of the
molecule through the membrane material. Separation of gas
components is therefore achieved by the differences in the
solubility and mobility behaviours of the components of the
natural gas feed.
Polymeric membranes can be further classied as rubbery
polymers and glassy polymers based on whether the membrane
process operates above or below the polymers glass transition
temperature. Glassy polymers operate below the glass transition
Table 9
Survey of membrane separation technologies used in the natural gas industry.
Vendor/
licensor
Membrane material Membrane module type Natural gas separation Flow rate
(MMscfd)
Refs.
Z-Top MTR Polymeric (Peruoro) Spiral-wound CO
2
from CH
4
to o2% 1300 www.mtrinc.com
LPG-Sep
TM
MTR Polymeric Spiral-wound LPG, NGLs from gas 250 www.mtrinc.com
Separex
TM
UOP Polymeric (cellulose acetate) Spiral-wound CO
2
from CH
4
, examples CO
2
in feed
reduced from 6.5% to 2%
Up to 680 www.uop.com
Medal Air Liquide Polymeric (polyimide) Hollow bre www.medal.airliquide.com
PRISM
s
APCI Polymeric (polysulfone) Hollow bre CO
2
reduced from 4.5% to, 2% o8 www.airproducts.com
CO
2
membrane UBE Polymeric (polyimide) Hollow bre CO
2
reduced from 9.6% to o2% 14100 www.ube.com
NitroSep
TM
MTR Polymeric Spiral wound N
2
from CH
4
to 4% N
2
in product gas 0.4100 www.mtrinc.com
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 141
temperature where the polymeric chains are rigid and separation
ability is primarily a function of the difference in kinetic diameter
of the gaseous species. Rubbery polymers by contrast operate
above the glass transition temperature where the polymer chains
are elastic and mobile. In this case diffusion is no longer hindered
by molecular size; rather it is the differences in the solubility of
the gases in the polymer that determines performance.
Membrane performance is chiey characterised by two para-
meters (1) permeability, which is a measure of the volume of gas
the membrane can process and (2) selectivity, which is a measure
of the membranes ability to separate components. Other factors
such as chemical, thermal and mechanical stability, fouling
tendencies, working lifetime, production costs and modularity
are also important to the success or otherwise of a membrane
unit operation. Permeability, or the permeability coefcient P
e
, is
dened as the transport ux per unit of cross membrane driving
force (pressure difference) (mol m m
2
s
1
Pa
1
). The total gas
ux, J, through the membrane can be calculated using Eq. (12),
which shows that the volume of gas the membrane can process
can be increased with a larger permeability coefcient, a larger
membrane area A, a decrease in membrane thickness l, and/or an
increase in the pressure gradient between the feed and permeate
streams Dp p
f
p
p
:
J
P
e
A p
f
p
p
_ _
l
QA p
f
p
p
_ _
12
The second equality in Eq. (12) introduces a new coefcient
also used to compare membrane performance known as per-
meance Q, which is dened as the permeability per unit mem-
brane thickness (mol m
2
s
1
Pa
1
). While the permeability is an
inherent property of the material and is commonly used in the
membrane literature to compare membrane performance for
different materials it is, however, calculated by rearranging Eq.
(12), which requires a knowledge or measurement of the mem-
brane thickness. For ultrathin membranes this is not always
possible and so to avoid confusion, the remainder of this article
will only discuss permeance data.
1
Gas transport of component i through the dense polymer
membranes utilised in natural gas processing can be further
described by (Baker and Lokhandwala, 2008)
J
i

D
i
K
i
A p
i,f
p
i,p
_ _
l
13
where J
i
is the volume (molar) ux (std cm
3
of component
i/(cm
2
s)) and l the membrane thickness. The diffusion coefcient
D
i
describes the motion of the gas molecule within the membrane
material, and this parameter is largely dependent on the size of
the gas molecule. The gas sorption coefcient K
i
is an indicator of
the solubility of the gas molecules in the membrane material.
The selectivity of a membrane is often represented in the
scientic literature as the permselectivity, or ideal selectivity,
determined by the permeance ratio of two components as given
by Eq. (14):
S
ij

Q
i
Q
j

D
i
D
j
_ _
K
j
K
i
_ _
14
Here Q
i
and Q
j
are the permeances of component i and j,
respectively, obtained from single component permeation
measurements. However, for real gas mixtures, such as natural
gas streams, the transport of gas molecules through the mem-
brane will be affected by the interactions between gas compo-
nents, as well as the interaction between the components and
membrane material. Hence, the true ability of a membrane to
separate components from a mixture deviates from the ideal
selectivity, and separation performance for mixtures should
instead be described by a separation coefcient S
ij
c
x
i
=x
j
_ _
=
y
i
=y
j
_ _
(Koros et al., 1996), which is analogous to the a
ij
dened
in Eq. (1). In this case, x
i,j
and y
i,j
are the mole fraction of
the components i and j in the permeate and feed streams,
respectively.
From these relationships it can be seen that the selectivity of
polymeric membranes is a function of the materials chemistry
and not the process parameters. In particular, the selectivity is a
function of the ratio of the diffusion coefcients (D
i
/D
j
) which is
proportional to the ratio of molecular diameters of the two
permeants, and the ratio of the solubility coefcients (K
i
/K
j
) of
the components in the membrane material. Furthermore, a trade-
off exists between a membranes permeance and selectivity
which is encapsulated by the empirical Robeson upper bounds
for gas pairs (Robeson, 1991, 2008). The Robeson upper bounds
highlight the maximum expected performance of rubbery and
glassy polymer membranes for gas separation applications. These
limits also provide a reference standard to which new membrane
materials can be compared. Many research groups have sought to
develop new materials that can push membrane performance for
gas separations beyond the current Robeson limits. Strategies for
improving membrane performance include the incorporation of
inorganic materials with high sorption capacities or molecular
sieving capabilities into a polymer matrix to create mixed matrix
membranes (Merkel et al., 2002) and synthesis of facilitated
transport membranes (FTMs). These novel membrane technolo-
gies will be further discussed relative to their specic applications
in natural gas processing.
Inorganic membranes including membranes prepared from
silica, zeolites and carbon molecular sieves have potential appli-
cations in natural gas processing. These porous inorganic mem-
branes function by a molecular sieving or activated transport
mechanism, in contrast to the sorption-diffusion mechanism
discussed above for polymeric membranes. Inorganic membranes
that can operate at high temperatures (500 1C and above), where
polymeric membranes are not stable, are the focus of signicant
research efforts aimed at developing technologies for post-com-
bustion CO
2
capture from ue gases (Duke et al., 2010), syngas
processing and H
2
purication (Lu et al., 2007; Smart et al., 2010;
Smith Scheinder et al., 2007). However, the removal of CO
2
and N
2
from natural gas does not require such high temperature pro-
cesses, and could in fact benet from sub-ambient process
temperatures, and so the use of porous inorganic membranes in
natural gas processing has not been widely reported, although the
separation of NGLs from CH
4
with zeolitic membranes has been
reported (Arruebo et al., 2001).
5.2. Membrane separation processes
Membranes operate as a semi-permeable barrier wherein a
component of the gas mixture will pass through the membrane
given a sufcient driving force or chemical potential gradient
between the feed and the permeate stream (Fig. 10). In the case of
natural gas processing the driving force takes the form of a partial
pressure gradient (Dp
i
p
f ,i
p
p,i
, for component i with partial
pressures p
f,i
and p
p,i
at the feed and permeate sides of the
membrane, respectively). The process and energy requirements
of a membrane system are limited to pre-treatment and compres-
sion of the feed stream to generate the desired driving force.
1
A barrer is a common unit of measure for permeability of a membrane
(10
-10
std cm
3
cm cm
2
s
1
cmHg
1
) whilst a GPU gas permeation unit is a unit
for permeance introduced in an attempt to simplify the units so that
1 GPU10
6
std cm
3
cm
2
s
1
cmHg
1
. To convert between the two units, one
needs to know the membrane thickness. As an example a membrane with a
permeability of 10 barrer and a thickness of 0.1 mm would have a permeance of
100 GPU.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 142
In the case of natural gas processing, the feed stream is often
already at sufcient pressure, although additional compression
may be required.
The separation ability of the membrane will determine whether
recycling and/or reprocessing of the permeate or retentate streams
in multiple stages is required to achieve the process objectives
(Fig. 10). Poor selectivity will dictate multiple membrane stages and
excessive recycling to attain the product purities required, increas-
ing the complexity of the process, energy usage, overall footprint of
the membrane unit operation and its capital cost.
The principal parameters in Eqs. (12) and (13) demonstrate
that the engineering design of a membrane system for CO
2
or N
2
removal from natural gas is determined by the volume of gas to
be processed, the feed gas pressure, concentration of the con-
taminants in the feed gas, and the required purity of the product
from the membrane unit. The membrane material has a large
effect on the capital cost of the membrane unit as the perfor-
mance of the membrane material determines both the membrane
area and the number of stages required. A signicant amount of
research has focussed on reducing the cost of membrane units
through the production of asymmetric composite membranes
which utilise an ultra-thin layer of high performance polymer
on a porous support as shown in Fig. 11. This serves the dual
purpose of reducing the quantity of the high performance poly-
mer and, as illustrated in Eq. (13), the ux increases because the
membrane is thinner. This strategy has been applied to both at
sheet and hollow bre geometries (Al-Juaied and Koros, 2006;
Omole et al., 2011; Strathmann, 2001).
The operational cost of a membrane unit is principally deter-
mined by the compression requirements; compression of the feed
gas and inter-stage streams is required to meet the partial
pressure gradients necessary to achieve sufcient component ux
across each membrane stage. The number of membrane stages
required is determined by the membrane selectivity for CO
2
/CH
4
or N
2
/CH
4
gas pairs and the desired level of contaminant removal
per stage. Most process simulation studies on membrane separa-
tion units for the removal of CO
2
from natural gas or CO
2
/CH
4
binary gas mixtures, report that a conguration of 2 or 3 mem-
brane stages is required to be most efcient and cost-effective
(Bhide and Stern, 1993a,b; Datta and Sen, 2006; Hao et al., 2002,
2008; Makaruk et al., 2010; Qi and Henson, 1998). Only once a
partial pressure gradient between the feed and the permeate has
been established can the membrane unit operate at all. Thus, if
the concentration of the component in the feed to be removed is
low, then the pressure of the feed stream must be increased
sufciently to establish the required driving force for sufcient
gas component ux through the membrane. For example, a
natural gas stream with 4% CO
2
would need to be compressed
to a pressure greater than 10,000 kPa to generate a permeate
stream of 80% CO
2
at 500 kPa. In addition, this partial pressure
gradient sets the product purity in the retentate stream; as once
the retentate and permeate streams reach equilibrium the net
ux through the membrane ceases. In the previous example, to
reduce the CO
2
content in the feed to below the required 50 ppm
for cryogenic gas processing, the membrane system would
require 10 stages with the highest pressure membrane unit
operating in excess of 9,830,000 kPa, assuming a 50% CO
2
stage
cut each time. Clearly, this simple analysis shows that membrane
separation technologies are more suited to bulk separation of CO
2
or N
2
from gas streams to meet pipeline specications than they
are to gas purication for cryogenic processing. To overcome
these limitations, hybrid membrane-amine absorption processes
have been designed, for example by UOP (UOP Overview of Gas
Processing Technologies and Applications, 2010), in which a
membrane separation unit rst provides a bulk removal of CO
2
from the sour gas and then a conventional amine absorption unit
is used to reduce the CO
2
content of the gas to less than 50 ppmv.
There are several other issues which must also be taken into
consideration when using membrane systems for the processing
of natural gas. Chief among these issues is the loss of selectivity
and separation performance that results from plasticisation and
degradation of polymer membranes during operation. Plasticisa-
tion can be caused by the common impurities found in natural gas
including CO
2
, water and aromatic hydrocarbons. When the
polymer matrix swells from plasticisation the increase in free
volume ultimately leads to a reduction in the relative differences
of the diffusion coefcients of the components to be separated,
and thus a deterioration in membrane selectivity is observed
(Koros and Mahajan, 2000; Visser et al., 2005; Xiao et al., 2009).
Glassy polymers suffer more severe performance degradation
than rubbery polymers and mixed matrix membranes. Although
sometimes plasticisation can be reversed, in many industrial
applications the regeneration of membranes is not feasible and
several membrane materials are known to undergo a swelling
induced conditioning history effect (Al-Juaied and Koros, 2006).
Membrane plasticisation can be managed by incorporating pre-
treatment operations to remove the contaminant plasticizers.
However, for membrane systems being used to remove CO
2
, the
plasticizing effect of the CO
2
cannot be mitigated by pre-treat-
ment (Ismail and Lorna, 2002, 2003; Visser et al., 2007). Pre-
treatment is also required to alleviate fouling from oil mists,
particulates and to prevent condensation of the heavier hydro-
carbons and water on the surface of the membranes as the phase
envelope for the natural gas often changes during operation
(Baker and Lokhandwala, 2008). Other strategies for minimising
plasticisation of polymer membranes include cross-linking of the
polymer network and/or the addition of llers to resist swelling
which have the added benet of increasing membrane selectivity
(Ismail and Lorna, 2002, 2003; Omole et al., 2011; Visser et al.,
2005; Visser et al., 2007; Xiao et al., 2009).
5.3. Application of membrane technology to CO
2
removal
Cellulose acetate (CA) is the earliest and the most commonly
applied polymeric membrane for CO
2
removal from natural gas.
Fig. 10. Block ow diagram for a single-stage membrane separation of CO
2
from
natural gas.
Fig. 11. Asymmetrical membrane morphology with dense selective layer on
porous support.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 143
Cellulose acetate is a glassy polymer that can be easily produced
as at sheets for spiral-wound membrane modules or as hollow
bres. Typically, CA displays permeance values of 60 GPU
(Zimmerman and Koros, 1999) and in laboratory tests has an
ideal CO
2
/CH
4
selectivity of 40. In eld trials the performance of
CA membranes is reduced to a selectivity of approximately 20
(Baker and Lokhandwala, 2008). Like all glassy polymers, CA
membranes are susceptible to plasticisation, especially from CO
2
and aromatic hydrocarbons. This is particularly problematic in
CO
2
removal applications given that membrane systems are most
suited to natural gas feeds containing high concentrations of CO
2
and yet these very same conditions can result in substantial
declines in selectivity (Donohue et al., 1989). By contrast, aro-
matic hydrocarbons need only be present in concentrations of
2001000 ppm to degrade the selectivity by 3050% (Vu et al.,
2002). However, CA membrane modules have also demonstrated
long term stability when exposed to high H
2
S concentrations (1
4% mol). For example, at the Kinder Morgan Yates gas plant the
CA membrane units have been reported to achieve normal work-
ing lifetimes of 3.5 yr despite increasing concentrations of CO
2
and H
2
S in the plant feed during this period (Kumar et al., 2011).
Despite the myriad of new polymers that have been developed
since the introduction of CA, few new materials have challenged
the widespread industrial application of CA membranes. Poly-
imides (PI) are temperature resistant polymers with a high glass
transition temperature and high ideal CO
2
/CH
4
selectivities of
between 3060 (Ayala et al., 2003; Stern, 1994). However PI
membranes, especially the uorinated PI polymers, are highly
susceptible to plasticisation in natural gas applications (Staudt-
Bickel and Koros, 1999). The other polymers of interest include
the polysulfones which demonstrate permeance values of 2040
barrer and CO
2
/CH
4
selectivities of 1535 (Gabelman et al., 2005)
and peruoro-type polymers which demonstrate increased resis-
tance to plasticisation in natural gas applications (Merkel et al.,
2006).
Permeation and selectivity data for a range of polymeric,
inorganic and mixed matrix membranes is presented in a mod-
ied Robeson plot shown in Fig. 12. This modied Robeson plot
takes account of the thickness of the membrane used in the
experiments, which allows direct comparison between polymeric
and inorganic membrane types of different thicknesses. Fig. 12
demonstrates that the industrial dominance of polymeric mem-
branes over their inorganic counterparts has less to do with
superior performance and more to do with the ease and econom-
ics of manufacturing membranes on a large scale.
Many of the synthesis and processing technologies for the
development of polymeric membranes for gas separation
undoubtedly grew from the earlier advances in membrane tech-
nology for water treatment (Reid and Breton, 1959). For example,
the successful methods for producing low cost polymeric mem-
branes in sufcient quantities to satisfy demand in the water
treatment industry has ensured that polymeric membranes have
been the main focus of membrane manufacturers for the last
three decades. Problems with plasticisation and low selectivities
have been, for the most part, overcome with clever process
engineering solutions. As a result, the drive for new materials
solutions has not been strong enough to warrant widespread
industrial investigation of inorganic membranes. Furthermore,
research groups specialising in inorganic membranes have
focussed specically on high temperature applications where
polymeric membranes are not feasible and have left the natural
gas processing space alone. Indeed, silica membranes are able to
achieve more than an order of magnitude higher permeance
compared to polymeric membranes with comparable selectivities.
However most of the results for silica membranes presented in
Fig. 12 are from studies where H
2
separation is the main focus and
the authors have investigated the behaviour of larger gas species
simply to probe the pore size distribution of the silica. Whether
industrial application of silica membranes for H
2
production is
feasible is still under investigation, and the effect of transfer of
any future developments in inorganic membrane technology from
H
2
purication to the natural gas industry remains difcult to
assess.
One solution that is gaining signicant attention in the
scientic literature is the incorporation of inorganic and polymer
materials into mixed matrix membranes (Merkel et al., 2002).
Fig. 12. Performance of various membranes for CO
2
/CH
4
separation. Polymer() Membranes refer to mixed matrix membranes. Robeson limit shown corresponds to 1 mm
membrane thickness. (Araki et al., 2007; Asaeda and Yamasaki, 2001; Battersby et al., 2006; Boffa et al., 2009; Boffa et al., 2008; de Lange et al., 1995a, b; de Vos et al., 1999; de
Vos and Verweij, 1998; Diniz da Costa et al., 2002; Diniz da Costa et al., 1999; Gopalakrishnan et al., 2006; Gu et al., 2008; Gu and Oyama, 2007; Gu and Oyama, 2009; Ikuhara
et al., 2007; Iwamoto et al., 2005; Jiang et al., 2004; Kim et al., 2001; Kim et al., 2003; Kusakabe et al., 1999; Kusakabe et al., 2003; Lee and Oyama, 2002; Lee et al., 2004; Li
et al., 2010a; Li et al., 2008; Li and Chung, 2008; Lokhandwala et al., 2010; Saimani et al., 2010; Syrtsova et al., 2004; Visser et al., 2005; Visser et al., 2007; Vu et al., 2002).
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 144
These combine the cheap, simple manufacturing techniques used
to produce polymeric membranes with the enhanced separation
characteristics of inorganic materials (Hu et al., 1997). In practice,
however, the synthesis of mixed matrix membranes is not a
simple task, with the compatibility of both phases crucial for
eliminating membrane defects and achieving good separation
performance (Mahajan and Koros, 2000; Zimmerman et al.,
1997). This is especially true for the glassy polymers employed
in natural gas processing, where poor interfacial interactions
between the polymer and the inorganic llers, for example
zeolites, degrade both membrane performance and mechanical
stability (Jiang et al., 2006). As a result, very few mixed matrix
membranes have displayed the performance enhancements initi-
ally promised, although some have broken Robesons upper
bound (Hillock et al., 2008).
Finally, and continuing the theme of selective enhancement,
some of the most exciting advances in membrane technology in
recent years have appeared in the form of facilitated transport
membranes (FTMs). These are membranes that incorporate mobile
carriers that selectively interact with CO
2
molecules to expedite the
ux of CO
2
through the membrane (Ebner and Ritter, 2009). The
selective interaction in FTMs can produce vastly superior membrane
performance compared to traditional polymeric membranes (Bara
et al., 2008; Moore et al., 2004). The four types of FTMs include thin
liquid lm membranes, ion-exchange membranes, polymer/metal
ion dispersions and modied polymer membranes. Thin liquid lm
membranes can be realised by depositing the liquid into a porous
support (immobilised liquid membrane), around bres (supported
liquid membrane) or between surfaces (contained liquid membrane)
(Bao and Trachtenberg, 2006). Chemicals used as the liquid mem-
brane include amine solutions, alkaline solutions such as potassium
carbonate, or an enzyme such as carbonic anhydrase which is
employed in the novel Carbozyme process (www.carbozyme.us).
The main obstacle to the successful deployment of FTMs in natural
gas processing applications are the well-publicised stability pro-
blems, in particular the evaporation of the carrier medium for
immobilised liquid membranes or water which acts as a swelling
medium in many ion-exchange membranes (Bara et al., 2010b;
Eriksen et al., 1993a, b; Noble et al., 1988). The evaporation of water
is less of a concern for natural gas feeds which, upstream of
dehydration units, may contain sufcient moisture to maintain
membrane performance (Ebner and Ritter, 2009).
The most common approach to overcome evaporation in
immobilised liquid membranes is to employ non-volatile solvents
such as polyelectrolytes and room temperature ionic liquids
(RTILs). These materials can also be considered to essentially
operate as high surface area liquid scrubbing units (Park et al.,
2009). Indeed, one particularly novel approach to the evaporation
problem is the so-called bulk-ow liquid membrane, where the
carrier medium is permeated and recycled through the support
pores in a manner resembling an ultra-high surface area scrub-
bing unit or membrane contactor (Duke et al., 2010). The most
promising FTMs are those with the carrier bonded to the polymer
backbone, which eliminates many of the stability concerns dis-
cussed above (Zhang et al., 2002). Ultimately, however, the
successful adoption of FTMs into natural gas processing will hinge
on the economics of producing large-scale membrane modules for
long-term operation, although the proposed trials of a FTM for
CO
2
separation in several large European power plants is highly
encouraging (Kim et al., 2004).
5.4. Application of membrane technology to N
2
rejection
In comparison to the extensive research and industrial work
undertaken using membranes for CO
2
removal from natural gas,
the use of membranes for N
2
removal is sparse. This is primarily
due to the competing and counterproductive separation coef-
cients of diffusion and solubility that render virtually all poly-
meric membranes unable to effectively remove either component.
The similar kinetic diameter of N
2
and CH
4
limits the effectiveness
of separation by diffusion and typical ideal N
2
/CH
4
selectivities for
glassy polymers are less than 3 and whilst CH
4
has a higher
solubility, ideal CH
4
/N
2
selectivities for rubbery polymers are less
than 5 (Lokhandwala et al., 2010). These values can be enhanced
slightly through the addition of inorganic llers to create mixed
matrix membranes. However, the separation behaviour of poly-
meric membranes for N
2
/CH
4
mixtures is so mediocre that no
Robeson upper bound currently exists. A comparison of N
2
/CH
4
membrane performance for polymeric and inorganic membranes
is shown in Fig. 13.
Fig. 13. Performance of various membranes for N
2
/CH
4
separation. Polymer() Membranes refer to mixed matrix membranes. Robeson limit shown corresponds to 1 mm
membrane thickness. (Araki et al., 2007; Asaeda and Yamasaki, 2001; Battersby et al., 2006; Boffa et al., 2009; Boffa et al., 2008; de Lange et al., 1995b; de Vos et al., 1999;
de Vos and Verweij, 1998; Gopalakrishnan et al., 2006; Ikuhara et al., 2007; Iwamoto et al., 2005; Kim et al., 2001; Kim et al., 2003; Kusakabe et al., 1999; Kusakabe et al.,
2003; Li et al., 2008; Li and Chung, 2008; Lokhandwala et al., 2010; Saimani et al., 2010; Syrtsova et al., 2004; Visser et al., 2005; Visser et al., 2007; Vu et al., 2002).
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 145
Fig. 13 indicates, however, the potential of inorganic membranes
for this natural gas processing application with silica achieving
selectivities of N
2
/CH
4
greater than 10, with permeances of more
than 2 orders of magnitude higher than comparable polymeric
membranes. Indeed, silica membranes can reach a performance
level that could only be afforded by theoretical polymer membranes
with a membrane thickness less than a single atom. However,
despite this potential, the difculties of N
2
removal from natural
gas streams are clearly illustrated in Fig. 13 in the slope of the
Robeson upper bounds. For every order of magnitude increase in
selectivity, permeance drops 4.5 orders of magnitude for N
2
per-
meation in Fig. 13, by comparison CO
2
permeation only drops
2.6 orders of magnitude. Despite this poor performance, membrane
systems for N
2
removal from natural gas have been realised
industrially through clever process engineering solutions. In parti-
cular they employ multi-stage or multi-step process designs to
theoretically treat a variety of feed streams with N
2
concentrations
varying from 4% to 30%. Engineering and economic analysis shows
that the cost of treatment rises sharply with increasing N
2
feed
concentrations and desired product purity. As a result most com-
mercial systems, of which NitroSep
TM
from Membrane Technology
and Research, Inc. are the most prominent, are for natural gas feeds
with less than 16% N
2
and reduce the N
2
concentration by roughly
half (Lokhandwala et al., 2010).
6. Separation by hydrates
6.1. Fundamentals of gas separation using hydrate phenomena
Clathrate hydrates are ice-like crystalline inclusion com-
pounds that consist of hydrogen-bonded water molecule cages,
which trap small molecules (o10

A) within their cages. Although
discovered in 1810 by Sir Humphrey Davy, it was not until
Hamerschmidt attributed problems in gas transport pipelines in
1934 to the formation of hydrates that signicant engineering
interest and research in these peculiar compounds took off.
Natural gas hydrates are now an important consideration for
most oil and gas production and processing activities due to their
tendency to agglomerate and cause a blockage in the pipeline or
process equipment (Sloan and Koh, 2008). Natural gas hydrates
also represent a potentially signicant energy resource with
current global estimates of (23) 10
3
gigatonnes of hydrate-
bound methane (Milkov, 2004).
There are three main structures for hydrates as illustrated in
Fig. 14a (Walsh, 2011). They consist of different congurations of
small (5
12
cages with 12 pentagonal faces) and large (5
12
6
2
,
5
12
6
4
, etc.) cages: for example structure I (sI) hydrate has two
small (5
12
) and six large (5
12
6
2
) cages. The equilibrium hydrate
structure (for simple hydrates) depends on the guest molecule as
illustrated in Fig. 14b. The structure is important for gas separa-
tion as certain molecules will only go into certain cages. For
example, propane can only enter the large cages of structure II
(sII) hydrate, whereas CH
4
and CO
2
can enter either cage.
Research on the intentional formation of gas hydrates has
mainly focussed on the use of hydrates as a medium for
transporting stranded gas in the form of natural gas hydrate
pellets (e.g., as an alternative to LNG). This concept was analysed
by Gudmundsson et al. (2004), who demonstrated the potentially
favourable economics of using gas hydrates over LNG for short to
medium transportation distances and smaller volumes of gas.
Hydrates have also been studied as a mechanism to store cool
energy for air-conditioning applications (Herri et al., 2005) and as
a method for desalination of seawater (Barduhn et al., 1962).
Using clathrate hydrates for gas separation has been explored
since the 1930s, where Nikitin used SO
2
hydrates to separate rare
gases (Sloan and Koh, 2008). Happel et al. (1994) show that
hydrates can be used to produce a gas stream leaner in the
hydrate forming gas. Upon dissociation of the hydrate, another
stream that is richer in the hydrate former is produced. They used
the rejection of N
2
from a gas mixture with CH
4
to demonstrate
this process. Lee and Kang (2003) showed that hydrates can be
used to separate CO
2
from ue gas. These studies lead the way for
various studies on the removal of CO
2
from ue gas (typically CO
2
and N
2
mixtures) using hydrates (Englezos et al., 2007b; Herri
et al., 2007; Kumar et al., 2006a; Lee et al., 2009) and also for the
pre-combustion removal of CO
2
from H
2
using hydrates (Englezos
et al., 2007a; Englezos et al., 2007b). The removal of CO
2
from
natural gas mixtures has received somewhat less attention (e.g.,
Golombok et al., 2009) as is the case for separating N
2
from a
natural gas mixtures (Happel et al., 1994; Hnatow and Happel,
1995; Johnson et al., 2009). However, at the recent triennial
international gas hydrate conference (ICGH7, July 1721 2011,
Edinburgh, UK) there were at least 28 papers submitted that
studied the separation of gases using hydrates.
The different afnity of certain gases to form hydrates (for
mixtures this results in a different composition of the gas in the
hydrate to the gas composition) allows the separation of mixture
Fig. 14. (a) Basic hydrate cage structures that combine to form the unit cells for the most common hydrate structures (Walsh, 2011), (b) hydrate structure versus size of
hydrate former (redrawn from Ripmeester, 2000).
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 146
components using hydrate formation. For CO
2
N
2
CH
4
gas mix-
tures, CO
2
is the more stable molecule in the hydrate phase,
whereas N
2
is the least stable. This results in the preferential
occupation of CO
2
over CH
4
in the hydrate cage, or in the absence
of CO
2
, the preferential occupation of CH
4
over N
2
. Seo et al.
(2001) showed that at lower pressures CO
2
also preferentially
occupies the hydrate cages from measurements of the composi-
tion of the gas phase and hydrate phase in equilibrium with a
CO
2
CH
4
gas mixture.
One of the limitations for the use of hydrates to separate gas
mixtures is the typically high pressure that is required to form gas
hydrates, which can represent a signicant cost to the process
(Tajima et al., 2004). A typical natural gas hydrate forms at high
pressure (430 bar) and low temperature (o20 1C). A common
method for overcoming this high pressure requirement is to use a
hydrate formation promoter molecule, which can signicantly
reduce the hydrate formation pressure for a given temperature. In
their patent, Lee and Kang (2003) in their patent proposed a
process for separating a multi-component gaseous mixture using
a hydrate promoter. The two most common examples of promo-
ter molecules that have been studied for hydrate separation
technologies include tetra-n-butylammonium bromide (TBAB)
(Clarke et al., 2011; Guo et al., 2011; Kamata et al., 2004; Li
et al., 2011; Li et al., 2010b; Shimada et al., 2003) and tetrahy-
drofuran (THF) (Lee and Kang, 2000; Lee et al., 2008; Sun et al.,
2010). Other promoters include cyclopentane (Fan et al., 2010)
and propane (Kumar et al., 2006b). It is important to note that the
addition of a promoter molecule can affect the selectivity of
certain gas molecules in the hydrate (Sun et al., 2011a) For
example, THF which occupies the large cage of sII hydrates has
been shown to selectively inhibit the enclathration of ethane,
which does not occupy the small cage at moderate pressures,
from a binary mixture with methane (Sun et al., 2011b).
6.2. Applications and demonstrations of hydrate gas separation for
CO
2
and N
2
capture.
There are a several patented processes proposed for gas
separation using hydrate formation although there is currently
no commercial-scale process in operation. Most of the processes
involve the use of a stirred tank reactor and include the use of
promoter molecules to reduce the operating pressure. An alter-
native approach is the use of a bubble-column to promote the
contact of gas and water (Chen et al., 2007), which is often a
limiting factor in the formation of gas hydrates.
The separation of CO
2
from a gas mixture using hydrates
typically requires the process to contain multiple steps to achieve
an adequate separation power because the composition of
enclatherated gas in the hydrate phase is often not very different
from that of the original gas mixture. For CO
2
/CH
4
separations,
van Denderen et al. (2009) reported values of a
ij
1.42.6 for
selectivity of the CO
2
in the hydrate phase. A typical apparatus
used to study the separation of a mixed gas using hydrates is
shown in Fig. 15 and described by Happel et al. (1994). The
apparatus produces a methane lean stream out of the hydrate
forming reactor (from a mixture of CH
4
N
2
) and methane rich
stream exiting the hydrate settler vessel. The apparatus would
likely represent a single stage in any separation process. The gas
stream exiting the hydrate forming reactor is lean in methane due
to the preferential occupancy of CH
4
in the hydrate cages over N
2
.
If the same apparatus were applied to treat a CO
2
CH
4
feed gas
mixture, the efuent from the top of the reactor would then be a
methane rich stream (lean in CO
2
) and the exit stream from the
settler vessel would be lean in methane (CO
2
rich).
Hydrate formation is an exothermic reaction and thus the
removal of heat from the process is an important consideration. A
novel approach of achieving this is the use of a uidised-bed heat
exchanger for hydrate formation (Waycuilis and York, 2002).
Marathon oil company has developed this method into a pilot-
scale continuous ow gas hydrate reactor capable of generating a
410 wt% hydrate slurry at 4750 kg/day production rate. Beads
of solid media are uidised inside the heat exchanger tubes by the
upward ow of the liquid which acts to continuously but gently
scour the inside tube surface. This helps to both (a) prevent solid
deposition from the hydrates, and (b) improve the internal heat
transfer coefcients. Preliminary engineering work has been
conducted to scale this process to a 3000 tonne/day slurry
generation (Waycuilis et al., 2011).
7. Summary and outlook
The characteristics of the primary gas separation operations
applied to CO
2
and N
2
removal from natural gas covered in this
review are summarised in Tables 10 and 11, respectively. For CO
2
removal, amine absorption technologies for acid gas treating are
well-established in the natural gas industry. The current industry
dominance of chemical absorption processes for acid gas treating
is likely to remain in the near future for the largest scale gas
processing applications because these processes have high
separation powers for CO
2
(and H
2
S) and the technology has been
proven. The use of CO
2
selective membrane technologies for bulk
separation of CO
2
is increasing in the natural gas industry and in
the capture of CO
2
from combustion ue gases (Olajire, 2010).
Novel low-temperature process technologies such as ExxonMo-
bils Controlled Freeze Zone
TM
(CFZ
TM
) process (Fieler et al., 2008)
Fig. 15. General schematic of an apparatus for producing a methane rich and a methane lean stream using hydrate formation and dissociation (redrawn from Happel et al.,
1994).
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 147
Table 10
Summary of characteristics for the main CO
2
removal (and gas sweetening) technologies.
Process technology Amines Physical solvents Hot potassium carbonate Condensation, distillation
or desublimation
Adsorption (PSA) Membranes
Mechanism/phase change Chemical absorption Absorption in liquid Chemical absorption Phase creation Selective CO
2
adsorption on solid
Permeation
Stage of deployment of
technology
Mature Mature Mature Commercial demo Commercial up to
2 MMscfd
Commercial
Commercial examples ADIP-X, Econamine
SM
,
aMDEA, GAS/SPEC
Selexol, purisol, rectisol Beneld, catacarb,
exsorb HP
Ryan/Holmes, CFZ
TM
,
Cryocell
s
Xebec PSA, Molecular
Gate
TM
Separex
TM
, Z-Top, Medal
CO
2
inlet concentration Up to 70% P
CO2
4 3.5 bar (GPSA
Engineering Data Book, 2004)
550% 3.565% Up to 40% Up to 90%
CO
2
outlet concentration 2%, down to 50 ppmv 50 ppmv possible Z1.5% (single-stage) 23%; 50 ppmv possible Down to 12%
50 ppmv in modied
CFZ
TM
(Nichols et al., 2009)
Z0.1% (two-stage)
(30 ppmv with hybrid)
(GPSA Engineering
Data Book, 2004)
Simultaneous H
2
S removal Yes (depends on solvent) Yes (most solvents) With two-stage scheme Yes, but limits on H
2
S in
CO
2
product
Possible Possible
Typical ow rate (MMscfd) Low to 4350 100400 Low to 4260 CDPs: 214 CDPs: Up to 2 Low to 4350
Typical operating conditions Pressure
(kPa)
Absorber: 50007000 Absorber: 65008000 Absorber: 50007000 20004000 10003500 200010,000
Regenerator: 150 Regenerator: 150
Temperature
(1C)
Absorber: 3060 73 to ambient 100116 142 to 45 25 o60 (materials limit)
Sweetened gas outlet pressure
(kPa)
50007000 65008000 50007000 20004000 10003500 200010000
Sweetened gas H
2
O saturation Outlet gas H
2
O saturated Gas dehydrated Outlet gas H
2
O saturated Gas dehydrated Gas dehydrated Gas dehydrated
Hydrocarbon recovery/losses
(Bergel and Tierno, 2009)
o1% losses Absorbs heavy hydrocarbons
and aromatics
Very low Very low Co-adsorption of
heavy hydrocarbons
1 stage: 815% (Bergel and
Tierno, 2009)
2 stages: o2%
Acid gas outlet pressure (kPa) 130 o500 Liquid CO
2
o500 o500
Footprint/layout considerations
(Bergel and Tierno, 2009)
High High High High; CFZ/CryocellLow Medium Low
Main equipment items (Bergel
and Tierno, 2009)
Contactor Contactor Contactor Cryogenic tower(s) Adsorbent vessels Membrane modules
Regenerator column Multiple ash drums Regeneration system Refrigeration system Waste/regen gas
compressors
Pretreatment modules
Flash tank Heat exchangers Gas/gas exchangers Gas/gas exchangers Valve and piping skids Compressors (2 stage
processes)
Lean/rich amine heat
exchanger
CO
2
ash drum Lean solution cooler Liquid CO
2
pump
Lean amine cooler Recycle compressors Circulation pumps
Solvent circulation
pumps
Chiller
Solvent circulation pumps
Energy requirements (main use) High (solvent
regeneration)
Medium (solvent circulation) Medium (regeneration) Medium (refrigeration
system)
Low (purge gas/
recompression)
Lowmedium (feed gas and
interstage compression)
Comparative process costs (Bergel
and Tierno, 2009)
Capital High Medium High Medium Medium Medium
Operating Medium Low Low Medium Low 1 stage: low
2 stages: medium
T
.
E
.
R
u
f
f
o
r
d
e
t
a
l
.
/
J
o
u
r
n
a
l
o
f
P
e
t
r
o
l
e
u
m
S
c
i
e
n
c
e
a
n
d
E
n
g
i
n
e
e
r
i
n
g
9
4
-
9
5
(
2
0
1
2
)
1
2
3

1
5
4
1
4
8
Table 11
Summary of characteristics for nitrogen rejection technologies.
Item Distillation Physical solvents Adsorption (PSA)N
2
selective Adsorption (PSA)CH
4
selective
Membranes
Mechanism/phase change Condensation Absorption in liquid
(typically CH
4
absorbs)
Selective N
2
adsorption on solid Selective CH
4
adsorption on
solid
Permeation
Commercial examples Linde, Praxair, Costain,
ConocoPhillips, APCI
AET; TGPE Molecular Gate
TM
Nitrotec; IAXC NitroSep
TM
Technology development
stage
Mature Mature Commercial up to 10 MMscfd Commercial up to 15 MMscfd Commercial up to 25 MMscfd
N
2
inlet concentration 855% Up to 40% Up to 40% Up to 40% Up to 90%
Typical ow rate (MMscfd) 15630 230 0.510 (Mitariten, 2009) 0.515 0.525 (Hale and Lokhandwala,
2004)
N
2
in CH
4
-rich product gas o1% o4% Typically 4% Typically 4% 1 stage bulk separation only
2 stages o4%
Hydrocarbon losses o2% Absorbs heavy
hydrocarbons
and aromatics
Propane and heavier adsorb on
binder of molecular sieve
(Mitariten, 2009)
1 stage: High
2 stages: low losses (o2%)
Typical operating conditions Pressure (kPa) 13002800 2003000 4002000 4002000 200010,000
Temperature (1C) 161 to 110 1C 32 to ambient 25 25 2560
Sales gas outlet pressure
(kPa)
2040 o5 4002000 o400 200010000
Footprint/ layout
comparative
High High Medium Medium Low
Main equipment items Cryogenic tower(s) Contactor Adsorbent vessels Adsorbent vessels Membrane modules
Refrigeration system Multiple ash drums
Gas/gas exchangers Recycle compressors
Solvent circulation pumps
Waste/regen gas compressors Compressors
(2 stage processes)
Pretreatment modules
Chiller Valve and piping skids
Energy requirements
(overall)
High (refrigeration
system)
Medium (solvent
circulation,
recompression)
Low Medium (CH
4
-rich gas
recompression)
Lowmedium (feed gas and
interstage compression)
Comparative process costs Capital Medium (with LNG) Medium Medium Medium Medium
Operating Medium Medium Low Low Medium
T
.
E
.
R
u
f
f
o
r
d
e
t
a
l
.
/
J
o
u
r
n
a
l
o
f
P
e
t
r
o
l
e
u
m
S
c
i
e
n
c
e
a
n
d
E
n
g
i
n
e
e
r
i
n
g
9
4
-
9
5
(
2
0
1
2
)
1
2
3

1
5
4
1
4
9
and adsorption-based systems such as Xebecs rapid cycle pres-
sure-swing adsorption (PSA) process (Toreja et al., 2011) are
emerging as alternatives to amine scrubbers for CO
2
removal in
certain applications such as for processing high CO
2
concentration
gases and for developing remote gas elds. Although the applica-
tion of hydrate formation to capture CO
2
is gathering signicant
research attention, with a large increase in the number of papers
on this topic observed at the recent 7th International Conference
on Gas Hydrates, the separation performance of hydrates is limited
by the relatively small differences between the free energies for
CO
2
and CH
4
in the hydrate cage. Thus, hydrates are unlikely to be
applied widely for treating sour natural gas elds.
From the technologies for N
2
rejection listed in Table 11,
cryogenic distillation is the leading NRU technology for large
scale (feed rates greater than 15 MMscfd) natural gas and LNG
plants, but technologies based on CH
4
selective absorption (AET-
Technology, 2007) and adsorption (Reinhold, 2010; Richter et al.,
1985) are commercially available for smaller scale gas processing
facilities. In the future there is potential for more efcient N
2
rejection technologies based on N
2
selective solvents and N
2
selective rapid cycle PSA systems. There remains signicant scope
for the development of better performing CO
2
selective mem-
branes, N
2
selective solvents and N
2
selective adsorbents to both
improve the separation power of, and the stability of materials, in
these emerging gas processing technologies.
Acknowledgements
We thank David Uhlmann (UQ), Hatim Essajee (UWA),
Mohamed El-Zaemey (UWA) and Thomas Saleman (UWA) for
their assistance with data collection during preparation of this
manuscript. The support of the Chevron Energy Technology
Company through the Chevron-UWA University Partnership Pro-
gram, the Western Australian Energy Research Alliance (WA:ERA)
and the Australian Research Council for UWAs gas process
engineering research program is also acknowledged.
References
Abanades, J.C., Rubin, E.S., Anthony, E.J., 2004. Sorbent cost and performance in
CO
2
capture systems. Ind. Eng. Chem. Res. 43, 34623466.
Ackley, M.W., Rege, S.U., Saxena, H., 2003. Application of natural zeolites in the
purication and separation of gases. Micropor. Mesopor. Mater. 61, 2542.
Ackley, M.W., Yang, R.T., 1991. Diffusion in ion-exchanged clinoptilolites. AlChE J.
37 (11), 16451656.
AET-Technology, 2007. 16 December 2010, /http://www.aet.com/tech.htmS.
Agrawal, R., Herron, D.M., Rowles, H.C., Kinard, G.E., 2003. Cryogenic Technology.
Kirk-Othmer Encyclopedia of Chemical Technology, vol. 8. John Wiley & Sons.
Al-Juaied, M., Koros, W.J., 2006. Performance of natural gas membranes in the
presence of heavy hydrocarbons. J. Membr. Sci. 274 (12), 227243.
Alcaniz-Monge, J., Lozano-Castello , D., Cazorla-Amoro s, D., Linares-Solano, A.,
2009. Fundamentals of methane adsorption in microporous carbons. Micropor.
Mesopor. Mater. 124, 110116.
Alcaniz-Monge, J., Marco-Lozar, J.P., Lillo-Ro denas, M.A., 2010. CO
2
separation by
carbon molecular sieve prepared from nitrated coal tar pitch. Fuel Process.
Technol. 92 (5), 915919.
Allen, A.D., Harris, R.O., Loescher, B.R., Stevens, J.R., Whiteley, R.N., 1973. Dinitro-
gen complexes of transition metals. Chem. Rev. 73, 1120.
Allen, A.D., Senoff, C.V., 1965. Nitrogenopentammineruthenium(II) complexes.
Chem. Commun., 621622.
Alonso-Vicario, A., Ochoa-Go mez, J.R., Gil-Ro, S., Go mez-Jime nez-Aberasturi, O.,
Ramrez-Lo pez, C.A., Torrecilla-Soria, J., Domnguez, A., 2010. Purication and
upgrading of biogas by pressure swing adsorption on synthetic and natural
zeolites. Micropor. Mesopor. Mater. 134, 100107.
Alvarado, D.B., Asaro, M., Bomben, J.L., Damle, A.S., Bhown, A.S., 1996. Nitrogen
Removal from Low Quality Gas. US Department of Energy, SRI International.
American Energies Pipeline, LLC, 2009. Improve Gas QualityNitrogen Rejection
Unit 29/10/2010, /http://www.nitrogenrejectionunit.com/facts.htmlS..
An, H., Feng, B., Su, S., 2011. CO
2
capture by electrothermal swing adsorption with
activated carbon bre materials. Int. J. Greenh. Gas Con. 5, 1625.
Araki, S., Mohri, N., Yoshimitsu, Y., Miyake, Y., 2007. Synthesis, characterization
and gas permeation properties of a silica membrane prepared by high-
pressure chemical vapor deposition. J. Membr. Sci. 290 (12), 138145.
Arruebo, M., Coronas, J., Mene dez, M., Santamaria, J., 2001. Separation of hydro-
carbons from natural gas using silicalite membranes. Sep. Purif. Technol. 25
(13), 275286.
Asaeda, M., Yamasaki, S., 2001. Separation of inorganic/organic gas mixtures by
porous silica membranes. Sep. Purif. Technol. 25 (13), 151159.
Ayala, D., Lozano, A.E., de Abajo, J., Garca-Perez, C., de la Campa, J.G., Peinemann,
K.V., Freeman, B.D., Prabhakar, R., 2003. Gas separation properties of aromatic
polyimides. J. Membr. Sci. 215 (12), 6173.
Bae, Y.-S., Hauser, B.G., Farha, O.K., Hupp, J.T., Snurr, R.Q., 2011. Enhancement of
CO
2
/CH
4
selectivity in metal-organic frameworks containing lithium cations.
Micropor. Mesopor. Mater. 141 (13), 231235.
Bae, Y.-S., Lee, C.-H., 2005. Sorption kinetics of eight gases on a carbon molecular
sieve at elevated pressure. Carbon 43 (1), 95107.
Bai, H., Yeh, A.C., 1997. Removal of CO
2
greenhouse gas by ammonia scrubbing.
Ind. Eng. Chem. Res. 36, 24902493.
Baker, R.W., Lokhandwala, K., 2008. Natural gas processing with membranes: an
overview. Ind. Eng. Chem. Res. 47, 21092121.
Bao, L., Trachtenberg, M.C., 2006. Facilitated transport of CO
2
across a liquid
membrane: comparing enzyme, amine and alkaline. J. Membr. Sci. 280, 330334.
Bao, Z., Alnemrat, S., Yu, L., Vasiliev, I., Ren, Q., Lu, X., Deng, S., 2011a. Kinetic
separation of carbon dioxide and methane on a copper metal-organic frame-
work. J. Colloid Interface Sci. 357, 504509.
Bao, Z.B., Yu, L.A., Ren, Q.L., Lu, X.Y., Deng, S.G., 2011b. Adsorption of CO
2
and CH
4
on a magnesium-based metal organic framework. J. Colloid Interface Sci. 353
(2), 549556.
Bara, J.E., Camper, D.E., Gabriel, C.J., Friedman, B.M., Israel, A., 2010a. Gas
processing with ionic liquid-amine solvents. In: 89th Annual Convention of
the Gas Processors Association, Austin, TX.
Bara, J.E., Camper, D.E., Gin, D.L., Noble, R.D., 2010b. Room-temperature ionic
liquids and composite materials: platform technologies for CO
2
capture. Acc.
Chem. Res. 43 (1), 152159.
Bara, J.E., Gabriel, C.J., Hatakeyama, E.S., Carlisle, T.K., Lessmann, S., Noble, R.D.,
Gin, D.L., 2008. Improving CO
2
selectivity in polymerized room-temperature
ionic liquid gas separation membranes through incorporation of polar sub-
stituents. J. Membr. Sci. 321 (1), 37.
Barduhn, A.J., Towlson, H.E., Hu, Y.C., 1962. The properties of some new gas
hydrates and their use in demineralizing sea water. AIChE J. 8 (2), 176183.
Bates, E.D., Mayton, R.D., Ntai, I., Davis Jr, J.H., 2002. CO
2
capture by a task-specic
ionic liquid. JACS 124 (6), 926927.
Battersby, S., Teixeira, P.W., Beltramini, J., Duke, M.C., Rudolph, V., Diniz da Costa,
J.C., 2006. An analysis of the Peclet and Damkohler numbers for dehydrogena-
tion reactions using molecular sieve silica (MSS) membrane reactors. Catal.
Today 116 (1), 1217.
Beck, J.S., Vartuli, J.C., Roth, W.J., Leonowicz, M.E., Kresge, C.T., Schmitt, K.D.,
Chu, C.T.-W., Olson, D.H., Sheppard, E.W., McCullen, S.B., Higgins, J.B., Schlen-
ker, J.L., 1992. A new family of mesoporous molecular sieves prepared with
liquid crystal templates. J. Am. Chem. Soc. 114 (27), 1083410843.
Belmabkhout, G., De Weireld, G., Fr ere, M., 2004. High-pressure adsorption
isotherms of N
2
, CH
4
, O
2
, and Ar on different carbonaceous adsorbents.
J. Chem. Eng. Data 49, 13791391.
Belmabkhout, Y., De Weireld, G., Sayari, A., 2009. Amine-bearing mesoporous silica
for CO
2
and H
2
S removal from natural gas and biogas. Langmuir 25,
1327513278.
Bergel, M. and I. Tierno, 2009. Sweetening technologiesa look at the whole
picture. In: 24th World Gas Conference. International Gas Union, Buenos Aires,
Argentina, pp. 117.
Bhide, B.D., Stern, S.A., 1993a. Membrane processes for the removal of acid gases
from natural gas. I. Process congurations and optimization of operating
conditions. J. Membr. Sci. 81 (3), 209237.
Bhide, B.D., Stern, S.A., 1993b. Membrane processes for the removal of acid gases
from natural gas. II. Effects of operating conditions, economic parameters, and
membrane properties. J. Membr. Sci. 81 (3), 239252.
Bindwal, A.B., Vaidya, P.D., Kenig, E.Y., 2011. Kinetics of carbon dioxide removal by
aqueous diamines. Chem. Eng. J. 169, 144150.
Bishnoi, S., Rochelle, G.T., 2002. Absorption of carbon dioxide in aqueous
piperazine/methyldiethanolamine. AlChE J. 48 (2), 27882799.
Boffa, V., ten Elshof, J.E., Garcia, R., Blank, D.H.A., 2009. Microporous niobia-silica
membranes: inuence of sol composition and structure on gas transport
properties. Micropor. Mesopor. Mater. 118 (13), 202209.
Boffa, V., ten Elshof, J.E., Petukhov, A.V., Blank, D.H.A., 2008. Microporous niobia-
silica membrane with very low CO
2
permeability. Chemsuschem 1 (5),
437443.
Bomberger, D.C., Bomben, J.L., Amirbahman, A., Asaro, M., 1999. Nitrogen Removal
from Natural Gas: Phase 2. US Department of Energy, SRI International,
Pittsburgh.
BP Statistical Review of World Energy June 2011 , 2011.
Breck, D.W., 1964. Crystalline molecular sieves. J. Chem. Educ. 41 (12), 678689.
Breck, D.W., 1974. Zeolite Molecular Sieves: Structure, Chemistry, and Use. Wiley,
New York.
Brennecke, J.F., Gurkan, B.E., 2010. Ionic liquids for CO
2
capture and emission
reduction. J. Phys. Chem. Lett. 1 (24), 34593464.
Burgers, W.F.J., Northrop, P.S., Kheshgi, H.S., Valencia, J.A., 2011. Worldwide
development potential for sour gas. Energy Procedia 4, 21782184.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 150
Burr, B., Lyddon, L., 2008. A Comparison of Physical Solvents for Acid Gas Removal.
Bryan Research & Engineering, Inc.
Butwell, K.F., W.B. Dolan and S.M. Kuznicki, 2002a. Pressure Swing Adsorption
Process. USA Patent Application US 6,497,750 B2.
Butwell, K.F., W.B. Dolan and S.M. Kuznicki, 2002b. Pressure Swing Adsorption
Process. USA patent application 6,497,750 B2.
Cai, J.J., Hawboldt, K., Abedinzadegan Abdi, M., 2012. Contaminant removal from
natural gas using dual hollow ber membrane contactors. J. Membr. Sci. 397398,
916.
Camper, D.E., Bara, J.E., Gin, D.L., Noble, R.D., 2008. Room-temperature ionic
liquidamine solutions: tunable solvents for efcient and reversible capture
of CO
2
. Ind. Eng. Chem. Res. 47, 84968498.
Cansado, I.P.P., Maur~ ao, P.A.M., Ribeiro Carrott, M.L., Carrott, P.J.M., 2010. Activated
carbons prepared from natural and synthetic raw materials with potential
applications in gas separations. Adv. Mater. Res. 107, 17.
Cavenati, S., Grande, C.A., Lopes, F.V.S., Rodrigues, A., 2009. Adsorption of small
molecules on alkali-earth modied titanosilicates. Micropor. Mesopor. Mater.
121, 114120.
Cavenati, S., Grande, C.A., Rodrigues, A., 2005. Separation of methane and nitrogen
by adsorption on carbon molecular sieve. Sep. Sci. Technol. 40, 27212743.
Cavenati, S., Grande, C.A., Rodrigues, A., 2006. Separation of CH
4
/CO
2
/N
2
mixtures
by layered pressure swing adsorption for upgrade of natural gas. Chem. Eng.
Sci. 61, 38933906.
Cavenati, S., Grande, C.A., Rodrigues, A.E., 2004. Adsorption equilibrium of
methane, carbon dioxide, and nitrogen on zeolite 13X at high pressures. J.
Chem. Eng. Data 49 (4), 10951101.
Chatt, J., Dillsworth, J.R., Richards, R.L., 1978. Recent advances in the chemistry of
nitrogen xation. Chem. Rev. 78 (6), 589625.
Chen, G.J., Luo, Y.T., Zhu, J.H., Fan, S.S., 2007. Study on the kinetics of hydrate
formation in a bubble column. Chem. Eng. Sci. 62 (4), 10001009.
Clarke, M.A., Acosta, H.Y., Bishnoi, P.R., 2011. Experimental measurements of the
thermodynamic equilibrium conditions of tetra-n-butylammonium bromide semi-
clathrates formed from synthetic landll gases. J. Chem. Eng. Data 56 (1), 6973.
Clodic, D., R. El Hitti, M. Younes, A. Bill and F. Casier, 2005. CO
2
capture by anti-
sublimation thermo-economic process evaluation. In: 4th Annual Conference
on Carbon Capture & Sequestration. Alexandria (VA), USA, .
CO
2
extraction & sequestration project Riley Ridge, W.Y., 2011. 21 June 2011,
/http://www.bcck.com/CO2-extraction-sequestration-projects/ultra-high-
CO2-removal-sequestration-plant-WY.htmlS.
CO2 Sponge, 2011. 19 July 2011, /http://www.iacx.com/co2sponge/S.
Collie, J., Hlavinka, M., Ashworth, A., 1998. An analysis of BTEX emissions from
amine sweetening and glycol dehydration facilities. 1998 Laurance Reid Gas
Conditioning Conference Proceedings. Norman, Oklahoma.
Contreras, M., Lagos, G., Escalona, N., Soto-Garrido, G., Radovic, L.R., Garcia, R., 2009. On
the methane adsorption capacity of activated carbons: in search of a correlation
with adsorbent properties. J. Chem. Technol. Biotechnol. 84, 17361741.
Controlled Freeze Zone
TM
increasing the supply of clean burning natural gas,
2010. /www.exxonmobil.comS. ExxonMobil.
Darde, V., Thomsen, K., van Well, W.J.M., Stenby, E.H., 2010. Chilled ammonia
process for CO
2
capture. Int. J. Greenh. Gas Con. 4, 131136.
Datta, A.K., Sen, P.K., 2006. Optimization of membrane unit for removing carbon
dioxide from natural gas. J. Membr. Sci. 283 (12), 291300.
de Lange, R.S.A., Hekkink, J.H.A., Keizer, K., Burggraaf, A.J., 1995a. Formation and
characterization of supported microporous ceramic membranes prepared by
sol-gel modication techniques. J. Membr. Sci. 99 (1), 5775.
de Lange, R.S.A., Hekkink, J.H.A., Keizer, K., Burggraaf, A.J., 1995b. Permeation and
separation studies on microporous sol-gel modied ceramic membranes.
Micropor. Mater. 4 (23), 169186.
de Vos, R.M., Maier, W.F., Verweij, H., 1999. Hydrophobic silica membranes for gas
separation. J. Membr. Sci. 158 (12), 277288.
de Vos, R.M., Verweij, H., 1998. High-selectivity, high-ux silica membranes for
gas separation. Science 279 (5357), 17101711.
Delgado, J., Uguina, M.A., Go mez, J.M., Ortega, L., 2006. Adsorption equilibrium of
carbon dioxide, methane and nitrogen onto Na- and H-mordenite at high
pressures. Sep. Purif. Technol. 48, 223228.
Dietzel, P.D.C., Besikiotis, V., Blom, R., 2009. Application of metal-organic frame-
works with coordinatively unsaturated metal sites in storage and separation of
methane and carbon dioxide. J. Mater. Chem. 19, 73627370.
Diniz da Costa, J.C., Lu, G.Q., Rudolph, V., Lin, Y.S., 2002. Novel molecular sieve
silica (MSS) membranes: characterisation and permeation of single-step and
two-step sol-gel membranes. J. Membr. Sci. 198 (1), 921.
Diniz da Costa, J.C., Lu, G.Q., Zhu, H.Y., Rudolph, V., 1999. Novel composite
membranes for gas separation: preparation and performance. J. Porous Mater.
6 (2), 143151.
Donohue, M.D., Minhas, B.S., Lee, S.Y., 1989. Permeation behavior of carbon
dioxide-methane mixtures in cellulose acetate membranes. J. Membr. Sci. 42
(3), 197214.
Drage, T.C., Arenilas, A., Smith, K.M., Pevida, C., Piippo, S., Snape, C.E., 2007.
Preparation of carbon dioxide adsorbents from the chemical activation of
ureaformaldehyde and melamineformaldehyde resins. Fuel 86, 2231.
Dreisbach, F., Staudt, R., Keller, J.U., 1999. High pressure adsorption data for
methane, nitrogen, carbon dioxide and their binary and ternary mixtures on
activated carbon. Adsorption 5, 215227.
Duke, M.C., Ladewig, B., Smart, S., Rudolph, V., da Costa, J.C.D., 2010. Assessment of
postcombustion carbon capture technologies for power generation. Front.
Chem. Eng. China 4 (2), 184195.
Dunne, J.A., Rao, M., Sircar, S., Gorte, R.J., Myers, A.L., 1996. Calorimetric heats of
adsorption and adsorption isotherms. 2. O
2
, N
2
, Ar, CH
4
, C
2
H
6
, and SF
6
on NaX,
H-ZSM-5, and Na-ZSM-5 zeolites. Langmuir 12, 58965904.
Ebner, A.D., Ritter, J.A., 2009. State-of-the-art adsorption and membrane separa-
tion processes for carbon dioxide production from carbon dioxide emitting
industries. Sep. Sci. Technol. 44, 12731421.
Eddaoudi, M., Li, H., Yaghi, O.M., 2000. Highly porous and stable metal-oxide
frameworks: structure design and sorption properties. J. Am. Chem. Soc. 122
(7), 13911397.
Englezos, P., Linga, P., Kumar, R., 2007a. The clathrate hydrate process for post and
pre-combustion capture of carbon dioxide. J Hazard Mater. 149 (3), 625629.
Englezos, P., Linga, P., Kumar, R.N., 2007b. Gas hydrate formation from hydrogen/
carbon dioxide and nitrogen/carbon dioxide gas mixtures. Chem. Eng. Sci. 62
(16), 42684276.
Eriksen, O.I., Aksnes, E., Dahl, I.M., 1993a. Facilitated transport of ethene through
Naon membranes. Part I. Water swollen membranes. J. Membr. Sci. 85 (1),
8997.
Eriksen, O.I., Aksnes, E., Dahl, I.M., 1993b. Facilitated transport of ethene through
Naon membranes. Part II. Glycerine treated, water swollen membranes. J.
Membr. Sci. 85 (1), 99106.
Esteves, I.A.A.C., Lopes, M.S.S., Nunes, P.M.C., Mota, J.P.B., 2008. Adsorption of
natural gas and biogas components on activated carbon. Sep. Purif. Technol.
62, 281296.
Faghihian, H., Pirouzi, M., 2009. Nitrogen separation from natural gas by modied
clinoptilolite. Clay Miner. 44, 289292.
Fan, S.S., Li, S.F., Wang, J.Q., Lang, X.M., Wang, Y.H., 2010. Clathrate hydrate capture
of CO(2) from simulated ue gas with cyclopentane/water emulsion. Chin. J.
Chem. Eng. 18 (2), 202206.
Favre, E., Svendsen, H.F., 2012. Membrane contactors for intensied post-combus-
tion carbon dioxide capture by gasliquid absorption processes. J. Membr. Sci.
407408, 17.
Fieler, E.R., Grave, E.J., Northrop P.S., Yeh, N.K., 2008. Controlled Freeze Zone
Tower. USA patent application WO 2008/091316. A1.
Figueroa, J.D., Fout, T., Plasynski, S., McIlvried, H., Srivastava, R.D., 2008. Advances
in CO
2
capture technologythe U.S. Department of Energys Carbon Sequestra-
tion Program. Int. J. Greenh. Gas Con. 2, 920.
Finn, A., 2007. Rejection strategies. Hydrocarbon Eng. 12 (10), 4952.
Friesen, D.T., Babcock, W.C., Edlund, D.J,. Lyon D.K., Miller, W.K., 2000. Liquid
absorbent solutions for separating nitrogen from natural gas. . USA patent
application 6,136,222. .
Friesen, D.T., Babcock, W.C., Edlund, D.J., Miller, W.K., 1993. Nitrogen sorption. USA
patent application 5,225,174.
Gabelman, A., Hwang, S.-T., Krantz, W.B., 2005. Dense gas extraction using a
hollow ber membrane contactor: experimental results versus model predic-
tions. J. Membr. Sci. 257 (12), 1136.
Gal, E., 2006. Ultra cleaning of combustion gas including the removal of CO
2
patent application WO/2006/022885. .
Golombok, M., van Denderen, M., Ineke, E., 2009. CO(2) removal from contami-
nated natural gas mixtures by hydrate formation. Ind. Eng. Chem. Res. 48 (12),
58025807.
Gonzalez-Garza, D., Rivera-Tinoco, R., Bouallou, C., 2009. Comparison of ammonia,
monoethanolamine, diethanolamine and methyldiethanolamine solvents to
reduce CO
2
greenhouse gas emissions. Chem. Eng. Trans. 18, 279284.
Gopalakrishnan, S., Nomura, M., Sugawara, T., Nakao, S.-i., 2006. Preparation of a
multi-membrane module for high-temperature hydrogen separation. Desali-
nation 193 (13), 230235.
GPSA Engineering Data Book, 2004. Gas Processors Suppliers Association, Tulsa,
OK.
Grande, C.A., Rodrigues, A., 2008. Electric swing adsorption for CO
2
removal from
ue gases. Int. J. Greenh. Gas Con. 2, 194202.
Gray, M.L., Soong, Y., Champagne, K.J., Pennline, H., Baltrus, J.P., Stevens Jr., R.W.,
Khatri, R., Chuang, S.S.C., Filburn, T., 2005. Improved immobilized carbon
dioxide capture sorbents. Fuel Process. Technol. 86 (1415), 14491455.
Gu, Y., Hacarlioglu, P., Oyama, S.T., 2008. Hydrothermally stable silica-alumina
composite membranes for hydrogen separation. J. Membr. Sci. 310 (12),
2837.
Gu, Y., Oyama, S.T., 2007. High molecular permeance in a poreless ceramic
membrane. Adv. Mater. 19 (12), 16361640.
Gu, Y., Oyama, S.T., 2009. Permeation properties and hydrothermal stability of
silica-titania membranes supported on porous alumina substrates. J. Membr.
Sci. 345 (12), 267275.
Gudmundsson, J.S., Parlaktuna, M., Khokhar, A.A., 2004. Storing natural gas as
frozen hydrate. SPE Prod. Facilities 9 (7), 6973.
Guild Associates, Inc., 2007. Molecular Gate Adsorption Technology 18 November
2010, /http://www.moleculargate.comS.
Guo, X.Q., Sun, Q.A., Liu, A.X., Liu, B., Huo, Y.S., Chen, G.Y., 2011. Experimental
study on the separation of CH(4) and N(2) via hydrate formation in TBAB
solution. Ind. Eng. Chem. Res. 50 (4), 22842288.
Gurkan, B.E., de la Fuente, J.C., Mindrup, E.M., Ficke, L.E., Goodrich, B.F., Price, E.A.,
Schneider, W., Brennecke, J.F., 2010. Equimolar CO
2
absorption by anion-
functionalized ionic liquids. JACS 132 (7), 21162117.
Habgood, H.W., 1958. Kinetics of molecular-sieve action. Sorption of nitrogen/
methane mixtures by Linde Molecular Sieve 4A. Can. J. Chem. 36, 13841397.
Hale, P., Lokhandwala, K., 2004. Advances in membrane materials provide new
solutions in the gas business. In: Laurance Reid Gas Conditioning Conference
2004., Oklahoma University of Oklahoma, Norman, p. 2.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 151
Hao, G.-P., Li, W.-C., Lu, A.-H., 2011. Novel porous solids for CO
2
capture. J. Mater.
Chem. 21, 64476451.
Hao, J., Rice, P.A., Stern, S.A., 2002. Upgrading low-quality natural gas with
H
2
S- and CO
2
-selective polymer membranes: Part I. Process design and
economics of membrane stages without recycle streams. J. Membr. Sci. 209 (1),
177206.
Hao, J., Rice, P.A., Stern, S.A., 2008. Upgrading low-quality natural gas with H
2
S-
and CO
2
-selective polymer membranes: Part II. Process design, economics, and
sensitivity study of membrane stages with recycle streams. J. Membr. Sci. 320
(12), 108122.
Happel, J., Hnatow, M.A., Meyer, H., 1994. The study of separation of nitrogen from
methane by hydrate formation using a novel apparatus. Int. Conf. Natural Gas
Hydrates 715, 412424.
Hart, A., Gnanendran, N., 2009. Cryogenic CO
2
capture in natural gas. Energy
Procedia 1, 697706.
Havran, V., Dudukovic, M.P., Lo, C.S., 2011. Conversion of methane and carbon
dioxide to higher value products. Ind. Eng. Chem. Res. 50, 70897100.
Heldebrant, D.J., Koech, P.K., Rainbolt, J.E., Zheng, F., 2011. CO
2
binding organic
liquids, an integrated acid gas capture system. Energy Procedia 4, 216223.
Herna ndez-Huesca, R., Diaz, L., Aguilar-Armenta, G., 1999. Adsorption equilibria
and kinetics of CO
2
, CH
4
and N
2
in natural zeolites. Sep. Purif. Technol. 15 (2),
163173.
Herri, J.M., Darbouret, M., Cournil, M., 2005. Rheological study of TBAB hydrate
slurries as secondary two-phase refrigerants. Int. J. Refrig. 28 (5), 663671.
Herri, J.M., Duc, N.H., Chauvy, F., 2007. CO
2
capture by hydrate crystallizationa
potential solution for gas emission of steelmaking industry. Energ. Convers.
Manage. 48 (4), 13131322.
Himeno, S., Komatsu, T., Fujita, S., 2005. High-Pressure adsorption equilibrua of
methane and carbon dioxide on several activated carbons. JCED 50 (2),
369376.
Hillock, A.M.W., Miller, S.J., Koros, W.J., 2008. Crosslinked mixed matrix mem-
branes for the purication of natural gas: effects of sieve surface modication.
J. Membr. Sci. 314 (12), 193199.
Hnatow, M.A., Happel, J., 1995. Process and apparatus for separation of constitu-
ents of gases using gas hydrates. USA patent application US 5,434,330.
Holmes, A.S., Ryan, J.M., 1982. Cryogenic distillative separation of acid gases from
methane. USA patent application US Patent 4,318,723.
Hu, Q., Marand, E., Dhingra, S., Fritsch, D., Wen, J., Wilkes, G., 1997. Poly(amide-
imide)/TiO
2
nano-composite gas separation membranes: fabrication and
characterization. J. Membr. Sci. 135 (1), 6579.
Huang, Z., Xu, L., Li, J.-H., Guo, G.-M., Wang, Y., 2009. Adsorption equilibrium of
carbon dioxide and methane on b-zeolite at pressures of up to 2000 kPa using
a static volumetric method. J. Chem. Eng. Data 55 (6), 21232127.
Hubbard, B., 2010. New and Emerging Technologies (Petroskills workshop). 2010
Gas Processors Association Convention. John M. Campbell & Co, Austin, Texas.
Hughes, T.J., Honari, A., Graham, B.F., Chauhan, A.S., Johns, M.L., May, E.F., 2012.
CO
2
sequestration for enhanced gas recovery: new measurements of super-
critical CO
2
CH
4
dispersion in porous media and a review of recent research.
Int. J. Greenh. Gas Con. 9, 457468.
Hughes, T.J., Syed, T., Graham, B.F., Marsh, K.N., May, E.F., 2011. Heat capacities and
low temperature thermal transitions of 1-hexyl and 1-octyl-3-methylimida-
zolium bis(triuoromethylsulfonyl)amide. J. Chem. Eng. Data 56 (5),
21532159.
Ikuhara, Y.H., Mori, H., Saito, T., Iwamoto, Y., 2007. High-temperature hydrogen
adsorption properties of precursor-derived nickel nanoparticle-dispersed
amorphous silica. J. Am. Ceram. Soc. 90 (2), 546552.
Ismail, A.F., Lorna, W., 2002. Penetrant-induced plasticization phenomenon in
glassy polymers for gas separation membrane. Sep. Purif. Technol. 27 (3),
173194.
Ismail, A.F., Lorna, W., 2003. Suppression of plasticization in polysulfone mem-
branes for gas separations by heat-treatment technique. Sep. Purif. Technol. 30
(1), 3746.
Iwamoto, Y., Sato, K., Kato, T., Inada, T., Kubo, Y., 2005. A hydrogen-permselective
amorphous silica membrane derived from polysilazane. J. Eur. Ceram. Soc. 25
(23), 257264.
Jayaraman, A., Herna ndez-Maldonado, A.J., Yang, R.T., Chinn, D., Munson, C.L.,
Mohr, D.H., 2004. Clinoptilolites for nitrogen/methane separation. Chem. Eng.
Sci. 59, 24072417.
Jensen, N.K., May, E.F., Watson, G., Rufford, TE., Zhang, D., and Chan, K.I. 2011.
Development of a method to efciently screen adsorbents for gas separation
applications involving methane, nitrogen, and carbon dioxide.. In: 9th Inter-
national Symposium on Characterisation of Porous Solids (IX COPS). Dreseden,
Germany.
Jensen, N.K., Rufford, T.E., Watson, G.C.Y., Zhang, D., Chan, K.I., May, E.F., 2012.
Screening zeolites for gas separation applications involving methane, nitrogen,
and carbon dioxide. JCED 57 (1), 106113.
Jessop, P.G., Heldebrant, D.J., Li, X.W., Eckert, C.A., Liotta, C.L., 2005. Green
chemistry: reversible nonpolar-to-polar solvent. Nature 436, 1102.
Jiang, L., Chung, T.-S., Li, D.F., Cao, C., Kulprathipanja, S., 2004. Fabrication of
matrimid/polyethersulfone dual-layer hollow ber membranes for gas separa-
tion. J. Membr. Sci. 240 (12), 91103.
Jiang, L.Y., Chung, T.S., Kulprathipanja, S., 2006. An investigation to revitalize the
separation performance of hollow bers with a thin mixed matrix composite
skin for gas separation. J. Membr. Sci. 276 (12), 113125.
Johnson, A., Osegovic, J.P., Max, M.D., 2009. Hydrate gas separationa new
technology for removing nitrogen from natural gas. GPA Convention.
Kamata, Y., Oyama, H., Shimada, W., Ebinuma, T., Takeya, S., Uchida, T., Nagao, J.,
Narita, H., 2004. Gas separation method using tetra-n-butyl ammonium
bromide semi-clathrate hydrate. Jpn. J. Appl. Phys. 1 43 (1), 362365.
Karadas, F., Atilhan, M., Aparicio, S., 2010. Review on the use of ionic liquids (ILs)
as alternative uids for CO
2
capture and natural gas sweetening. Energy Fuels
24, 58175828.
Kelley, B.T., Valencia, J.A., Northrop, P.S., Mart, C.J., 2011. Controlled Freeze Zone
TM
for developing sour gas reserves. Energy Procedia 4, 824829.
Kerry, F.G., 2007. Industrial Gas Handbook: Gas Separation and Purication. CRC
Press, Boca Raton, Florida.
Kidnay, A.J., Parrish, W., 2006. Fundamentals of Natural Gas Processing. CRC Press,
Boca Raton.
Kim, T.J., Li, B.A., Hagg, M.B., 2004. Novel xed-site-carrier polyvinylamine
membrane for carbon dioxide capture. J. Polym. Sci. B-Polym. Phys. 42 (23),
43264336.
Kim, Y.J., You, J.K., Hong, W.H., Yi, K.B., Ko, C.H., Kim, J.-N., 2008. Characteristics of
CO
2
absorption into aqueous ammonia. Sep. Sci. Technol. 43, 766777.
Kim, Y.S., Kusakabe, K., Morooka, S., Yang, S.M., 2001. Preparation of microporous
silica membranes for gas separation. Kor. J. Chem. Eng. 18 (1), 106112.
Kim, Y.S., Kusakabe, K., Yang, S.M., 2003. Microporous silica membrane
synthesized on an ordered mesoporous silica sublayer. Chem. Mater. 15 (3),
612615.
Kluson, P., Scaife, S., Quirke, N., 2000. The design of microporous graphitic
adsorbents for selective separation of gases. Sep. Purif. Technol. 20, 1524.
Kohl, A., Nielsen, R., 1997. Gas Purication. Gulf Publishing Company, Houston,
Texas.
Koros, W.J., Ma, Y.H., Shimidzu, T., 1996. Terminology for membranes and
membrane processes. Pure Appl. Chem. 68 (7), 14791489.
Koros, W.J., Mahajan, R., 2000. Pushing the limits on possibilities for large scale gas
separation: which strategies. J. Membr. Sci. 175, 181196.
Kumar, M., Glass, J., Echt., W., Dyszkiewicz, C., 2011. Robust performance of
membranes in EOR gas processing with high H2S. In: Gas Processors Associa-
tion Annual Conference. San Antonio, Texas,pp. 19.
Kumar, R., Linga P., Englezos, P., 2006a. Pre and post combustion capture of carbon
dioxide via hydrate formationIn: EIC Climate Change Conference, vols. 1 and 2
: pp. 294-300.
Kumar, R., Wu, H.J., Englezos, P., 2006b. Incipient hydrate phase equilibrium for
gas mixtures containing hydrogen, carbon dioxide and propane. Fluid Phase
Equilibr. 244 (2), 167171.
Kusakabe, K., Sakamoto, S., Saie, T., Morooka, S., 1999. Pore structure of silica
membranes formed by a sol-gel technique using tetraethoxysilane and
alkyltriethoxysilanes. Sep. Purif. Technol. 16 (2), 139146.
Kusakabe, K., Shibao, F., Zhao, G., Sotowa, K.I., Watanabe, K., Saito, T., 2003. Surface
modication of silica membranes in a tubular-type module. J. Membr. Sci. 215
(12), 321326.
Kuznicki, S.M., 1990. Preparation of small-pored crystalline titanium molecular
sieve zeolite. USA patent application US 4,938,939.
Kuznicki, S.M., Bell, V.A., Nair, S., Hillhouse, H.W., Jacubinas, R.M., Braunbarth,
C.M., Toby, B.H., Tsapatsis, M., 2001. A titanosilicate molecular sieve with
adjustable pores for size-selective adsorption of molecules. Nature 412 (6848),
720724.
Kuznicki, S.M., Bell, V.A., Petrovic I., Blosser, P.W., 1999. Separation of nitrogen
from mixtures thereof woth methane utilizing barium exchanged ETS-4. USA
patent application US5,989,316.
Kuznicki, S.M., Bell, V.A., Petrovic, I., and Desai, B.T., 2000. Small-pored crystalline
titanium molecular sieve zeolites and their use in gas separation processes.
USA patent application US 6,068,682.
Larue, J., Lebas, E., 1996. Integrated process for natural gas treatment. Oil Gas Sci.
Technol. 51 (5), 653668.
Laverty, B., OHair, G., 1990. The applications of membrane technology in the
natural gas industry. In: Howell, J.A. (Ed.), The Membrane Alternative: Energy
Implications for Industry. Spon Press, pp. 176.
Lee, D., Oyama, S.T., 2002. Gas permeation characteristics of a hydrogen selective
supported silica membrane. J. Membr. Sci. 210 (2), PII S0376-
7388(0302)00389-00387.
Lee, D., Zhang, L., Oyama, S.T., Niu, S., Saraf, R.F., 2004. Synthesis, characterization,
and gas permeation properties of a hydrogen permeable silica membrane
supported on porous alumina. J. Membr. Sci. 231 (12), 117126.
Lee, H., Kang, S.P., 2000. Recovery of CO
2
from ue gas using gas hydrate:
thermodynamic verication through phase equilibrium measurements.
Environ. Sci. Technol. 34 (20), 43974400.
Lee, H., Kang, S.P., 2003. Method for separation of gas constituents employing
hydrate promoter. Patent application US 6.602,326 B2.
Lee, H., Seo, Y., Kang, S.P., Lee, S., 2008. Experimental measurements of hydrate
phase equilibria for carbon dioxide in the presence of THF, propylene oxide,
and 1,4-dioxane. J. Chem. Eng. Data 53 (12), 28332837.
Lee, J.W., Zhang, J.S., Yedlapalli, P., 2009. Thermodynamic analysis of hydrate-
based pre-combustion capture of CO(2). Chem. Eng. Sci. 64 (22), 47324736.
Lemmon, E.W., Huber, M.L., McLinden, M.O., 2010. REFPROP. NIST National
Institute of Standards and Technology,: Reference Fluid Thermodynamic and
Transport Properties.
Li, F., Li, Y., Chung, T.-S., Kawi, S., 2010a. Facilitated transport by hybrid POSS
s
-
Matrimid
s
-Zn2 nanocomposite membranes for the separation of natural
gas. J. Membr. Sci. 356 (12), 1421.
Li, J.-R., Kuppler, R.J., Zhou, H.-C., 2009. Selective gas adsorption and separation in
metal-organic frameworks. Chem. Soc. Rev. 38 (5), 14771504.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 152
Li, X.S., Xia, Z.M., Chen, Z.Y., Wu, H.J., 2011. Precombustion capture of carbon
dioxide and hydrogen with a one-stage hydrate/membrane process in the
presence of tetra-n-butylammoniurn bromide (TBAB). Energy Fuel 25 (3),
13021309.
Li, X.S., Xu, C.G., Chen, Z.Y., Wu, H.J., 2010b. Tetra-n-butyl ammonium bromide
semi-clathrate hydrate process for post-combustion capture of carbon dioxide
in the presence of dodecyl trimethyl ammonium chloride. Energy 35 (9),
39023908.
Li, Y., Chung, T.-S., Xiao, Y., 2008. Superior gas separation performance of dual-
layer hollow ber membranes with an ultrathin dense-selective layer. J.
Membr. Sci. 325 (1), 2327.
Li, Y., Chung, T.S., 2008. Highly selective sulfonated polyethersulfone (SPES)-based
membranes with transition metal counterions for hydrogen recovery and
natural gas separation. J. Membr. Sci. 308 (12), 128135.
Linstrom, P.J., Mallard, W.G. (Eds.), 2011. NIST Chemistry WebBook, NIST Standard
Reference Database Number 69. National Institute of Standards and Technol-
ogy, Gaithersburg, MD /http://webbook.nist.govS. (retrieved 24 August
2011).
Liu, W., King, D., Liu, J., Johnson, B., Wang, Y., Yang, Z., 2009. Critical material and
process issues for CO
2
separation from coal-powered plants JOM.
Lokhandwala, K.A., Pinnau, I., He, Z., Amo, K.D., DaCosta, A.R., Wijmans, J.G., Baker,
R.W., 2010. Membrane separation of nitrogen from natural gas: A case study
from membrane synthesis to commercial deployment. J. Membr. Sci. 346 (2),
270279.
Lu, G.Q., Diniz da Costa, J.C., Duke, M., Giessler, K., Socolow, R., Williams, R.H.,
Kreutz, T., 2007. Inorganic membranes for hydrogen production and purica-
tion: a critical review and perspective. J. Colloid Interface Sci. 314, 589603.
MacDowell, N., Florin, N., Buchard, A., Hallett, J., Galindo, A., Jackson, G., Adjiman,
C.S., Williams, C.K., Shah, N., Fennell, P., 2010. An overview of CO
2
capture
technologies. Energy Environ. Sci. 3, 16451669.
MacKenzie, D., Cheta, I., Burns, D., 2002. Removing nitrogen. Hydrocarbon Eng. 7
(11), 5763.
Mahajan, R., Koros, W.J., 2000. Factors controlling successful formation of mixed-
matrix gas separation materials. Ind. Eng. Chem. Res. 39 (8), 26922696.
Makaruk, A., Miltner, M., Harasek, M., 2010. Membrane biogas upgrading pro-
cesses for the production of natural gas substitute. Sep. Purif. Technol. 74 (1),
8392.
Mallick, A., Saha, S., Pachfule, P., Roy, S., Banerjee, R., 2010. Selective CO
2
and H
2
adsorption in a chiral magnesium-based metal organic framework (MgMOF)
with open metal sites. J. Mater. Chem. 20, 90739080.
Mansourizadeh, A., Ismail, A.F., 2009. Hollow ber gas-liquid membrane contac-
tors for acid gas capture: a review. J. Hazard. Mater. 171 (13), 3853.
Marathe, R.P., Mantri, K., Srinivasan, M.P., Farooq, S., 2004a. Effect of ion exchange
and dehydration temperature on the adsorption and diffusion of gases in ETS-
4. Ind. Eng. Chem. Res. 43 (17), 52815290.
Marathe, R.P., Mantri, K., Srinivasan, M.P., Farooq, S., 2004b. Effect of ion exchange
and dehydration temperature on the adsorption and diffusion of gases in ETS-
4. Ind. Eng. Chem. Res. 43 (17), 52815290.
McCabe, W.L., Smith, J.C., Harriott, P., 2005. Unit Operations of Chemical Engineer-
ing. McGraw-Hill, New York.
McCabe, W.L., Thiele, E.W., 1925. Graphical design of fractionating columns. Ind.
Eng. Chem. 17 (6), 605611.
Mehra, Y.R., Gaskin, T.K., 1997. Guidelines offered for choosing cryogenics or
absorption for gas processing. Oil Gas J. 97 (99), 6269.
Mehrotra, A., Ebner, A.D., Ritter, J.A., 2011. Simplied graphical approcah for
complex PSA cycle scheduling. Adsorption 17 (2), 337345.
Membrane Technology & Research, 2010. Nitrogen removal from natural gas:
Nitrosep 16 December 2010, /http://www.mtrinc.com/nitrogen_removal.htmlS.
Merkel, T.C., Freeman, B.D., Spontak, R.J., He, Z., Pinnau, I., Meakin, P., Hill, A.J.,
2002. Ultrapermeable, reverse-selective nanocomposite membranes. Science
296 (5567), 519522.
Merkel, T.C., Pinnau, I., He, Z., Prabhakar, R., Freeman, B.D., 2006. Gas and vapor
transport properties of peruoropolymers. In: Yampolskii, Y., Pinnau, I., Free-
man, B.D. (Eds.), Materials Science of Membranes for Gas and Vapor Separa-
tion. Wiley, Chichester, England.
Milkov, A.V., 2004. Global estimates of hydrate-bound gas in marine sediments:
how much is really out there? Earth-Sci Rev. 66 (34), 183197.
Miller, E.W., Soychak, S.J., Bartoo, R.K., Reed, A.E., Ackman, R., 1999. Brady plant
treating project. In: 1999 Laurance Reid Gas Conditioning Conference, 1999.
University of Oklahoma, Norman, OK.
Miller, W.K., Gilbertson, J.D., Leiva-Paredes, C., Bernatis, P.R., Weakley, T.J.R., Lyon,
D.K., Tyler, D.R., 2002. Inorg. Chem. 41, 54535465.
Mimura, T., Shimojo, S., Suda, T., Iijima, M., Mitsuoka, S., 1995. Research and
development on energy saving technology for ue gas carbon dioxide recovery
and steam system in power plant. Energy Convers. Manage. 36 (69),
397400.
Mitariten, M., 2009. Nitrogen removal from natural gas with the molecular gate
TM
adsorption process. In: 88th Annual Convention of the Gas Processors
Association 2009, 811 March 2009 San Antonio, Texas. Gas Processors
Association, pp. 544555.
Moore, T.T., Mahajan, R., Vu, D.Q., Koros, W.J., 2004. Hybrid membrane materials
comprising organic polymers with rigid dispersed phases. AIChE J. 50 (2),
311321.
Morrow, D.C. Lunsford, K.M., 1997. Removal and disposal of BTEX components
from amine acid gas streams. In: Seventy-Sixth GPA Annual Convention, Gas
Processors Association, Tulsa, Oklahoma., pp. 171173.
Nam, G.-M., Jeong, B.-M., Kang, S.-H., Lee, B.-K., Choi, D.-K., 2005. Equilibrium
isotherms of CH
4
, C
2
H
6
, C
2
H
4
, N
2
, and H
2
on zeolite 5A using a static volumetic
method. J. Chem. Eng. Data 50, 7276.
Nhattacharya, S., Watney, L., Newell, D., Magnuson, M., 2009. Demonstration of a
Low-cost 2-Tower Micro Scale N
2
Rejection System to Upgrade Low-BTU Gas
from Stripper Wells, Kansas Gelogical Society, University of Kansas Research
Center.
Nichols, J.L.V., Friedman, B.M., Nold, A.L., McCutcheon, S., Goethe, A., 2009.
Processing technologies for CO
2
rich gas. In: 88th Annual Convention of the
Gas Processors Association 2009, Gas Processors Association, San Antonio,
Texas.
Noble, R.D., Pellegrino, J.J., Grosgogeat, E., Sperry, D., Way, J.D., 1988. CO
2
separation using facilitated transport ion-exchange membranes. Sep. Sci.
Technol. 23 (1213), 15951609.
Normand, L., Streicher, C., Laborie, G., Lemaire, E., Raynal, L., 2012. Development of
technologies for CO
2
capture from ue gases. In: 91st Annual Gas Processors
Association Convention. New Orleans, Louisiana.
Olajire, A.A., 2010. CO
2
capture and sequestration technologies for end-of-pipe
applications. Energy 35, 26102628.
Olajossy, A., Gawdzik, A., Budner, Z., Dula, J., 2003. Methane separation from coal
mine methane gas by vacuum pressure swing adsorption. Chem. Eng. Res. Des.
81 (4), 474482.
Omole, I.C., Bhandari, D.A., Miller, S.J., Koros, W.J., 2011. Toluene impurity effects
on CO
2
separation using a hollow ber membrane for natural gas. J. Membr.
Sci. 369 (12), 490498.
Park, Y.-I., Kim, B.-S., Byun, Y.-H., Lee, S.-H., Lee, E.-W., Lee, J.-M., 2009. Preparation
of supported ionic liquid membranes (SILMs) for the removal of acidic gases
from crude natural gas. Desalination 236 (13), 342348.
Penny, D.E., Ritter, T.J., 1983. Kinetic study of the reaction between carbon dioxide
and primary amines. J. Chem. Soc., Faraday Trans. 1 79, 21032109.
Phan, L., Chiu, D., Heldebrant, D.J., Huttenhower, H., John, E., Li, X.W., Pollet, P.,
Wang, R.Y., Eckert, C.A., Liotta, C.L., Jessop, P.G., 2008. Switchable solvents
consisting of amidine/alcohol or guanidine/alcohol mixtures. Ind. Eng. Chem.
Res. 47 (3), 539545.
Plasynski, S.I., Litynski, J.T., McIlvried, H., Srivastava, R.D., 2009. Progres and new
developments in carbon capture and storage. Crit. Rev. Plant Sci. 28, 123128.
Plaza, M.G., Rubiera, F., Pis, J.J., Pevida, C., 2010. Ammoxidation of carbon materials
for CO
2
capture. Appl. Surf. Sci. 256, 68436849.
Plechkova, N.V., Seddon, K.R., 2008. Applications of ionic liquids in the chemical
industry. Chem. Soc. Rev. 37, 123150.
Qi, R., Henson, M.A., 1998. Optimal design of spiral-wound membrane networks
for gas separations. J. Membr. Sci. 148 (1), 7189.
Radebaugh, R., 2007. Historical summary of cryogenic activity prior to 1950. In:
Timmerhaus, K.D., Reed, R. (Eds.), Cryogenic Engineering. Springer.
Rajani, J.B., 2004. Treating technologies of shell global solutions for natural gas and
renery gas streams. In: Research Institue of Petroleum Industry Congress 12 Iran.
Reid, C.E., Breton, E.J., 1959. Water and ion ow across cellulosic membranes. J.
Appl. Polym. Sci. 1 (2), 133143.
Reinhold, H.E.I., 2010. Pressure Swing Adsorption Method And System Separating
Gas Components. USA Patent Application US7740687B2.
Rezaei, F., Webley, P.A., 2010. Structured adsorbents in gas separation processes.
Sep. Purif. Technol. 70, 243256.
Ribeiro, R.P., Sauer, T., Lopes, F.V.S., Moreira, R.F., Grande, C.A., Rodrigues, A.E.,
2008. Adsorption of CO
2
, CH
4
, and N
2
in activated carbon honeycomb
monolith. J. Chem. Eng. Data 53, 23112317.
Richter, E., Korbacher, W., Knoblauch, K., Giessler, K., 1985. Method of producing
methane-rich gas mixture from mine gas. USA patent application US
4521221.
Ripmeester, J.A., 2000. Hydrate research - from correlations to a knowledge-based
discipline: the importance of structure. In: 3rd International Conference on
Gas Hydrates, Salt Lake City, Utah.
Ritter, J.A., Ebner, A.D., 2007. State-of-the-art adsorption and membrane separa-
tion processes for hydrogen production in the chemical and petrochemical
industries. Sep. Sci. Technol. 42, 11231193.
Ritter, J.A., Yang, R.T., 1987. Equilibrium adsorption of multicomponent gas
mixtures at elevated pressures. Ind. Eng. Chem. Res. 26 (8), 16791686.
Robeson, L.M., 1991. Correlation of separation factor versus permeability for
polymeric membranes. J. Membr. Sci. 62 (2), 165185.
Robeson, L.M., 2008. The upper bound revisited. J. Membr. Sci. 320 (12), 390400.
Rojey, A., Jaffret, C., Cornot-Grandolphe, S., Durand, B., Jullian, S., Valais, M., 1997.
Natural Gas Production Processing Technology. Editions Technip, Paris.
Ruthven, D.M., 2011. Molecular sieve separations. Chem. Ing. Tech. 83 (12), 4452.
Saha, D., Bao, Z., Jia, F., Deng, S., 2010. Adsorption of CO
2
, CH
4
, N
2
O, and N
2
on
MOF-5, MOF-177, and Zeolite 5A. Environ. Sci. Technol. 44, 18201826.
Saimani, S., Dal-Cin, M.M., Kumar, A., Kingston, D.M., 2010. Separation perfor-
mance of asymmetric membranes based on PEGDa/PEI semi-interpenetrating
polymer network in pure and binary gas mixtures of CO
2
, N
2
and CH
4
. J.
Membr. Sci. 362 (12), 353359.
Scaife, S., Kluson, P., Quirke, N., 2000. Characterization of porous materials by gas
adsorption: do different molecular probes give different pore structures? J.
Phys. Chem. B 104, 313318.
Seader, J.D., Henley, E.J., 2006. Separation Process Principles. John Wiley & Sons,
Hoboken, USA.
Seagraves, J., Weiland, R., 2011. Troubleshooting amine plants using mass transfer
rate-based simulation tools. In: Laurance Reid Gas Conditioning Conference.
University of Oklahoma, Norman, Oklahoma.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 153
Sellmann, D., 1974. Dinitrogentransition metal complexes: synthesis, properties,
and signicance. Angew. Chem., Int. Edn. (Engl.) 13, 639649.
Seo, Y.T., Lee, H., Yoon, J.H., 2001. Hydrate phase equilibria of the carbon dioxide,
methane, and water system. J. Chem. Eng. Data 46 (2), 381384.
Sevilla, M., Valle-Vigo n, P., Fuertes, A.B., 2011. N-doped polypyrrole-based porous
carbons for CO
2
capture. Adv. Funct. Mater. 21 (14), 27812787.
Sheikh, M.A., Hassan, M.M., Loughlin, K.F., 1996. Adsorption equilibria and rate
parameters for nitrogen and methane on Maxsorb activated carbon. Gas Sep.
Purif. 10 (3), 161168.
Shimada, W., Ebinuma, T., Oyama, H., Kamata, Y., Takeya, S., Uchida, T., Nagao, J.,
Narita, H., 2003. Separation of gas molecule using tetra-n-butyl ammonium
bromide semi-clathrate hydrate crystals. Jpn. J. Appl. Phys. 2 42 (2A),
L129L131.
Skarstrom, C.W., 1960. Method and apparatus for fractionating gaseous mixtures
by adsorption. United States patent application US Patent 2944627.
Sloan, E.D., Koh, C.A., 2008. Clathrate Hydrates of Natural Gas. CRC Press, Taylor &
Francis Group.
Smart, S., Lin, C.X.C., Ding, L., Thambimuthu, K., Diniz da Costa, J.C., 2010. Ceramic
membranes for gas processing in coal gasication. Energy Environ. Sci. 3 (3),
268278.
Smith Scheinder, P., Duke, M., Liu, S., Abdel-jawad, M., Diniz da Costa, J.C., 2007.
Energetics for gas separation in microporous membranes. Int. J. Nanotechnol. 4
(5), 468481.
Song, C., 2006. Global challenges and strategies for control, conversion and
utilization of CO
2
for sustainable development involving energy, catalysis,
adsorption and chemical processing. Catal. Today 115, 232.
Staudt-Bickel, C., Koros, W.J., 1999. Improvement of CO
2
/CH
4
separation charac-
teristics of polyimides by chemical crosslinking. J. Membr. Sci. 155 (1),
145154.
Stern, S.A., 1994. Polymers for gas separations: the next decade. J. Membr. Sci. 94
(1), 165.
Strathmann, H., 2001. Membrane separation processes: current relevance and
future opportunities. AIChE J. 47 (5), 10771087.
Sun, C., Li, W., Yang, X., Li, F., Yuan, Q., Mu, L., Chen, J., Liu, B., Chen, G., 2011a.
Progress in research of gas hydrate. Chin. J. Chem. Eng. 19 (1), 151162.
Sun, C.Y., Chen, G.J., Zhang, L.W., 2010. Hydrate phase equilibrium and structure
for (methane plus ethane plus tetrahydrofuran plus water) system. J. Chem.
Thermodyn. 42 (9), 11731179.
Sun, Q., Guo, X., Liu, A., Liu, B., Huo, Y., Chen, G., 2011b. Experimental study on CH
4
and N
2
separation via hydrate formation in TBAB solution. Ind. Eng. Chem. Res.
50 (4), 22842288.
Syrtsova, D.A., Kharitonov, A.P., Teplyakov, V.V., Koops, G.H., 2004. Improving gas
separation properties of polymeric membranes based on glassy polymers by
gas phase uorination. Desalination 163 (13), 273279.
Tagliabue, M., Farrusseng, D., Valencia, S., Aguado, S., Ravon, U., Rizzo, C., Corma,
A., Mirodatos, C., 2009. Natural gas treating by selective adsorption: material
science and chemical engineering interplay. Chem. Eng. J. 155, 553566.
Tajima, H., Yamasaki, A., Kiyono, F., 2004. Energy consumption estimation for
greenhouse gas separation processes by clathrate hydrate formation. Energy
29 (11), 17131729.
TGPE, 2009. Nitrogen Rejection Units 11-11-2010, /www.tuckergas.comS.
Thiruvenkatachari, R., Su, S., Yu, X.X., 2009. Carbon bre composite for ventilation
air methane (VAM) capture. J. Hazard. Mater. 172, 15051511.
Thote, J.A., Iyer, K.S., Chatti, R., Labhsetwar, N.K., Biniwale, R.B., Rayalu, S.S., 2010.
In situ nitrogen enriched carbon for carbon dioxide capture. Carbon 48,
396402.
Todd, R.S., 2003. Experimental study of a rapid pressure swing adsorption system
for air separation. Ph.D. Thesis, Monash University.
Tonkovich, A.L., 2004. Upgrading methane using ultra-fast thermal swing adsorp-
tion, Velocys, U. DOE.
Toreja, J., Chan, N., VanNostrand, B., Dickinson, J.P., 2011. Rotary-valve, fast-cycle
pressure swing adsorption technology allows West Coast platofrm to meet
tight California specications and recover stranded gas. In: Laurance Reid Gas
Conditioning Conference. Norman, Oklahoma, University of Oklahoma.
UOP, 2010. Hydrogen Management with PSA and Recovery Equipment 11-11-
2010, /www.uop.com/rening/1100.htmlS.
UOP Overview of Gas Processing Technologies and Applications, 2010. 22 June
2011, /http://www.uop.com/wp-content/uploads/2011/02/UOP-Overvie
w-of-Gas-Processing-Technologies-and-Applications-tech-presentation.pdfS.
US Environmental Protection Agency, 2008. Upgrading Drained Coal Mine
Methane to Pipeline Quality: A Report on the Commercial Status of System
Suppliers. US Environmental Protection Agency, C. M. O. Program.
US EPA, 1997. Technical and Economic Assessment of Potential to Upgrade Gob
Gas to Pipeline Qualtiy. US Environmental Protection Agency, C. M. O.
Program.
Vaidya, P.D., Kenig, E.Y., 2007. CO
2
-alkanoamine reaction kinetics: a review of
recent studies. Chem. Eng. Technol. 30 (11), 14671474.
Valencia, J.A., Denton, R.D., 1983. Method and apparatus for separating carbon
dioxide and other acid gases from methane by the use of distillation and a
controlled freezing zone. USA patent application US Patent 4,533,372.
Valenzuela, D.P., Myers, A.L., 1989. Adsorption Equilibrium Data Handbook.
Prentice Hall, Engleword Cliffs, NJ.
Vargas, D.P., Giraldo, L., Silvestre-Albero, J., Moreno-Piraja n, J.C., 2011. CO
2
adsorption on binderless activated carbon monolith. Adsorption 17 (3),
497504.
Versteeg, G.F., Van Dijck, L.A.J., Van Swaaij, W.P.M., 1996. On the kinetics between
CO
2
and alkanoamines both in aqueous and non-aqueous solutions. an over-
view. Chem. Eng. Commun. 144 (1), 113158.
Visser, T., Koops, G.H., Wessling, M., 2005. On the subtle balance between
competitive sorption and plasticization effects in asymmetric hollow ber
gas separation membranes. J. Membr. Sci. 252 (12), 265277.
Visser, T., Masetto, N., Wessling, M., 2007. Materials dependence of mixed gas
plasticization behavior in asymmetric membranes. J. Membr. Sci. 306 (12),
1628.
Vu, D.Q., Koros, W.J., Miller, S.J., 2002. High pressure CO
2
/CH
4
separation using
carbon molecular sieve hollow ber membranes. Ind. Eng. Chem. Res. 41 (3),
367380.
Walsh, M., 2011. Molecular dynamics simulations of hydrates (needs xing). PhD,
Colorado School of Mines.
Wang, B., C ote , A.P., Furukawa, H., OKeefe, M., Yaghi, O.M., 2008. Colossal cages in
zeolitic imidazolate frameworks as selective carbon dioxide reservoirs. Nature
453, 207211.
Warmuzinski, K., 2008. Harnessing methane emissions from coal mining. Process
Saf. Environ. Prot. 86, 315320.
Watson, G., May, E.F., Graham, B.F., Trebble, M.A., Trengove, R.D., Chan, K.I., 2009.
Equilibrium adsorption measurement of pure nitrogen, carbon dioxide, and
methane on a carbon molecular sieve at cryogenic temperatures and high
pressures. J. Chem. Eng. Data 54, 27012707.
Watson, G.C.Y., Jensen, N.K., Rufford, T.E., Chan, K.I., May, E.F., 2012. Volumetric
adsorption measurements of N
2
, CO
2
, CH
4
, and a CO
2
CH
4
mixture on a
natural chabazite from 5 to 3000 kPa. J. Chem. Eng. Data 57 (1), 93101.
Waycuilis, J.J., Sirajuddin, H., Rensing, P., 2011. A novel approach to production of
hydrate slurries at high process intensity and conceptual applications. In: 7th
International Conference on Gas Hydrates, July 1721, Edinburgh.
Waycuilis, J.J., York, S.D., 2002. Production of a Gas Hydrate Slurry Using A
Fluidized Bed Heat Exchanger. Patent application US 6,350,928 B1.
Weiland, R., Hatcher, N.A., Nava, J.L., 2010. Benchmarking solvents for carbon
capture. In: A.B. de Haan, H. Kooijman and A. Go rak, (Eds.) Distillation &
Absorption, 1215 September 2010 Eindhoven, The Netherlands.
Werner, S., Marco, H., Wasserscheid, P., 2010. Ionic liquids in chemical engineer-
ing. Annu. Rev. Chem. Biomol. Eng. 1, 203230.
Wilkinson, D., Johnson, G., 2010. Nitrogen Rejection Technologies for Abu Dhabi.
5 August 2011, /www.nitrogen-rejection.comS.
Willems, G.P., Golombok, M., Tesselaar, G., Brouwers, J.J.H., 2010. Condensed
rotational separation of CO
2
from natural gas. AlChE J. 56 (1), 150159.
Wills, J., Mitariten, M., 2009. Production of Pipeline Quality Natural Gas With the
Molecular Gate
TM
CO
2
Removal Process. Tidelands Oil Production Company &
Guild Associates.
Xiao-Zhong, C., Yi-Jiang, Z., Yu-He, K., Wei-Guang, Z., Shou-Yong, Z., Ya-Ping, Z., Li,
Z., 2009. Dynamic experiments and model of hydrogen and deuterium
separation with micropore molecular sieve Y at 77K. Chem. Eng. J. 152,
428433.
Xiao, Y., Low, B.T., Hosseini, S.S., Chung, T.S., Paul, D.R., 2009. The strategies of
molecular architecture and modication of polyimide-based membranes for
CO
2
removal from natural gasa review. Prog. Polym. Sci. 34 (6), 561580.
Xu, X., Song, C., Andresen, J.M., Miller, B.G., Scaroni, A.W., 2002. Novel polyehty-
lenimine-modied mesoporous molecular sieve of MCM-41 Type as high-
capacity adsorbent for CO
2
capture. Energy Fuels 16 (6), 14631469.
Xu, X., Zhao, X., Sun, L., Liu, X., 2008. Adsorption separation of carbon dioxide,
methane, and nitrogen on Hb and Na-exchange b-zeolite. J. Natural Gas Chem.
17, 391396.
Zangeneh, F.T., Sahebdelfar, S., Ravanchi, M.T., 2011. Conversion of carbon dioxide
to valuable petrochemicals: an approach to clean development mechanism. J.
Natural Gas Chem. 20 (3), 219231.
Zhang, J., Webley, P.A., 2008. Cycle development and design for CO
2
capture from
ue gas by vacuum swing adsorption. Environ. Sci. Technol. 42, 563569.
Zhang, Y., Wang, Z., Wang, S.C., 2002. Selective permeation of CO
2
through new
facilitated transport membranes. Desalination 145 (13), 385388.
Zhang, Y., Zhang, S., Lu, X., Zhou, Q., Fan, W., Zhang, X., 2009. Dual amino-
functionalised phosphonium ionic liquids for CO
2
capture. Chem. Eur. J. 15
(12), 30033011.
Zhao, D., Feng, J., Huo, Q., Melosh, N., Fredrickson, G.H., Chmelka, B.F., Stucky, G.D.,
1998. Triblock copolymer syntheses of mesoporous silica with periodic 50 to
300 angstrom pores. Science 279 (5350), 548552.
Zhao, H., 2006. Innovative applications of ionic liquids as green engineering
liquids. Chem. Eng. Commun. 193, 16601677.
Zhou, S.J. Meyer,, H., Bikson, B., Ding, Y., 2010. Hybrid membrane absorption
process for post combustion CO
2
capture. 2010 AIChE Spring Meeting and 6th
Global Congress on Process Safety. San Antonio, TX.
Zimmerman, C.M., Koros, W.J., 1999. Comparison of gas transport and sorption in
the ladder polymer BBL and some semi-ladder polymers. Polymer 40 (20),
56555664.
Zimmerman, C.M., Singh, A., Koros, W.J., 1997. Tailoring mixed matrix composite
membranes for gas separations. J. Membr. Sci. 137 (12), 145154.
T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94-95 (2012) 123154 154

S-ar putea să vă placă și