The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

download The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

of 32

Transcript of The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    1/32

    The removal of CO2  and N2   from natural gas: A review of conventional andemerging 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, Australiab The University of Queensland, FIMLab—Films 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 2011Accepted 2 June 2012Available online 19 June 2012

    Keywords:

    CO2  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 CO2

    and N2  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 flow schemes designed to produce

    pipeline quality gas (typically 2% CO2, o3% N2) and gas to feed a cryogenic gas plant (typically 50 ppmv

    CO2, 1% N2). Amine absorption technologies for CO2   and H2S 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 CO2   selective

    membrane technologies for bulk separation of CO2 is increasing in the natural gas industry. Novel low-

    temperature CO2   removal technologies such as ExxonMobil’s Controlled Freeze ZoneTM process and

    rapid cycle pressure swing adsorption processes are also emerging as alternatives to amine scrubbers in

    certain applications such as for processing high CO2 concentration gases and for developing remote gas

    fields. Cryogenic distillation remains the leading N2   rejection technology for large scale (feed rates

    greater than 15 MMscfd) natural gas and liquefied natural gas plants. However, technologies based on

    CH4   selective absorption and adsorption, as well as N2  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 CO2 selective membranes, N2 selective solvents and N2selective 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 world’s 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 thedevelopment potential of unconventional, stranded and contami-

    nated gas reserves that were previously considered economically

    unviable. Furthermore, many of the significant natural gas reserves

    are located far from the large established gas markets in Western

    Europe, Japan and South Korea. Therefore, significant volumes of 

    natural gas must be transported long distances from exporting

    countries either by pipeline or by tanker as liquefied 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-Pacific region.

    The estimated world volumes of sub-quality natural gasreserves, including sour natural gas reserves, are significant.

    Sub-quality natural gas reserves are defined as gas fields contain-

    ing more than 2% CO2, 4 % N2   and 4 parts per million (ppm)

    hydrogen sulphide (H2S) (Kidnay and Parrish, 2006). Burgers et al.

    (2011), for example, estimate that 50% of the volume of known

    gas resources contain more than 2% CO2. 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 efficient

    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:  w ww.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:  [email protected] (E.F. May).

     Journal of Petroleum Science and Engineering 94 –95 (2012) 123–154

    http://www.elsevier.com/locate/petrolhttp://www.elsevier.com/locate/petrolhttp://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.petrol.2012.06.016mailto:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.petrol.2012.06.016http://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.petrol.2012.06.016mailto:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.petrol.2012.06.016http://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.petrol.2012.06.016http://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.petrol.2012.06.016http://www.elsevier.com/locate/petrolhttp://www.elsevier.com/locate/petrol

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    2/32

    natural gas. Additional gas processing challenges arise from new

    environmental regulations that may call for capture and seques-

    tration of CO2   from gas fields and stricter regulation of CH4emissions in N2   vent streams from natural gas production

    facilities.

    Carbon dioxide, as well as H2S and other acid gases, must be

    removed from natural gas because in the presence of water theseimpurities can form acids that corrode pipelines and other

    equipment. Although this paper is not focussed on the removal

    of H2S, 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 CO2   capture technologies described are more

    selective for H2S than CO2. Furthermore, CO2 provides no heating

    value and must be removed to meet gas quality specifications

    before distribution to gas users. The maximum level of CO2permitted in natural gas transmitted to customers by pipeline is

    typically less than 3% (Hubbard, 2010) although local contracts

    may stipulate quality specifications different to these values. The

    specifications of CO2 removal from natural gas to be processed in

    a cryogenic plant to produce LNG, recover liquefied petroleum gas

    (LPG) or natural gas liquids (NGL) are more stringent than thosefor typical gas pipelines. For example, in addition to more

    extensive dehydration, the CO2   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 specification for N2   is 3% (Kidnay and

    Parrish, 2006); because N2 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% N2 is often specified for

    LNG, for example to avoid stratification and rollover of the liquid

    product during shipping. To ensure this specification is met in

    plants liquefying NG with a high N2 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 N2   concentrations in natural gas

    processed by an LNG plant are energetically parasitic because a

    significant amount of energy is wasted cooling the N2   from

    ambient temperatures to those of LNG (about  161 1C at atmo-

    spheric pressure). Furthermore, even if the NG does not have a

    high N2   concentration, an NRU may be necessary because the

    liquefaction process produces two product streams: the LNG and

    an ‘end-flash’ gas which is a mixture of N2  and CH4  vapour. This‘end-flash’ 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-flash 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 CO2   and N2  removal

    from natural gas, with particular attention paid to large scale LNG

    production. The introduction sections provide an overview of a

    typical process flow 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 CO2 capture

    from CH4 can also be applied to CO2 capture from combustion flue

    gases, for example in coal-fired power stations. The recent

    advances and future trends in technologies for capturing CO2  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 flue gas (predo-

    minantly a mixture of CO2, N2 and H2O) treatment. The two most

    significant differences between these applications are the CO2partial pressure, and the level of CO2 removal required. In the first

    case, the natural gas feed to the CO2   removal unit is typically

    available at high pressures (more than 3000 kPa) while flue gases

    are typically at close to atmospheric pressure, so the driving force

    for CO2  capture from flue gas (partial pressure CO2  typically less

    than 15 kPa) is much lower than from natural gas. In the second

    case the level of CO2 removal required for natural gas production

    is greater, especially for LNG production plants, than bulk separa-

    tion of CO2   from flue gas. Thus, some promising strategies for

    CO2   capture in thermal power plants such as oxy-combustion

    (a modified combustion process that can produce a high CO2concentration flue gas, Plasynski et al., 2009) are not relevant to

    CO2  removal from natural gas.

    1.1. Overview of conventional gas processing flow 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

    specifications and the gas flow rate, the units identified 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 N2  concentrations, or if a high N2content gas can be blended with richer natural gas streams to

    meet pipeline gas specifications.

     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 4.7 100.63 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 75.35 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) 123–154124

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    3/32

    In conventional gas processing H2S and CO2  are removed in an

    acid gas removal unit (AGRU) using aqueous amine absorptionprocesses. The sweetened gas leaving the amine process is saturated

    with water, so typically the AGRU is located upstream of the

    dehydration facilities. The final 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 CO2, components that could freeze in

    the cryogenic plant include water (typically removed to less than

    0.1 ppmv), heavier paraffinic hydrocarbons and aromatics such as

    benzene. To protect the aluminium plate-fin heat exchangers used

    in the NRU the feed gas to the cryogenic plant must be free of 

    mercury (below 0.01 mg/Nm3)  (Kidnay and Parrish, 2006). The gas

    feed entering the (Pre-Cool section of the) liquefaction plant istypically 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 N2 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 fluid refrigeration cycles. If an NRU is incorporated

    with the liquefaction process then it would commonly be located

    downstream of the MCHE and before the final 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 CO2, N2 and

    CH4   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 purification 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) permeationthrough a membrane and (5) chemical conversion to another

    compound (Kohl and Nielsen, 1997;   Seader and Henley, 2006).

    The first four of these operations are discussed in this review;

    direct chemical conversion of CO2   from natural gas to a useful

    product is beyond the current scope. An example of such direct

    conversion, which is attracting significant current research inter-

    est, is the dry reforming process where CO2  reacts with CH4   to

    produce syngas (a mixture of H2 and CO2) which can then be used

    for the production of liquid fuels via Fischer–Tropsch reactions.

    Recent reviews on the conversion of CO2   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 liquefied natural gas production. The required operations and

    process flow arrangements vary depending on the local feed gas properties, size of the plant, and product specifications; and may not all be required (MCHE¼main

    cryogenic heat exchanger).

     Table 2

    Physical properties of methane, carbon dioxide and nitrogen.

    Property CH4   CO2   N2

    Kinetic diameter ( Å) (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 ( Å3) (Tagliabue et al., 2009)   2.448 2.507 1.710

    Quadrapole moment (D Å) (Tagliabue et al., 2009)   0.02 4.3 1.54

    T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94–95 (2012) 123–154   125

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    4/32

    desublimation of CO2. If the volatility differences among species

    to be separated are not sufficiently large, as is the case for

    N2–CH4, 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 separationmechanisms have more than one type of selectivity; however, the

    most common and applicable type is the equilibrium selectivity, aij,which is defined in terms of phase compositions:

    aij   xi x j

      y j yi

      ð1Þ

    Here  xi and x j are the mole fractions (or convenient concentra-

    tion units) of components in one equilibrium phase, and y i and  y jare the mole fraction of components in a second equilibrium

    phase. This definition 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 vapour–liquid

    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 entireseparation process in terms of the product stream compositions. For

    a single stage equilibrium operation aij ¼ SP ij; however, for a processutilizing multiple separation stages SP ij values much larger than the

    single stage equilibrium selectivity can be achieved.

    SP ij C ð1Þ

    i

    C ð1Þ j

    0@

    1A C ð2Þ j

    C ð2Þi

    !  ð2Þ

    Here C ð1Þi   and C ð2Þi   are the concentrations of component i in each of 

    the product streams, and C ð1Þ j   , C ð2Þ

     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  aij41 could be engineered to achieve a

    given separation power if no constraints exist on capital, operationaland energy costs. In practice a separation process will be selected if it

    can achieve the separation power necessary to meet the desired

    product specifications at a cost that is (i) lower than other processing

    alternatives, and (ii) economically viable in terms of the

    product’s value.

    In Table 3, indicative values of  aij and SP ij are given for the mostcommon industrial processes used to separate CO2 or N2 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 defined. For CO2removal from natural gas, a convenient approach may be to compare

    the aCO2,CH4  values of the various separation technologies by definingthe equilibrium phases to be used with Eq. (1) as the CO2  selective

    phase: that is the amine solution or physical solvent for absorption,

    the high-purity liquid CO2 product for distillation, the adsorbed phase

    ( xi or  qi) for gas–solid adsorption systems, and the permeate stream

    from a CO2  selective membrane stage. Using available data for CO2removal 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% CO2) from a feed mixture containing

    5% CO2   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 ijbaij  through the arrangement of multipleseparation stages into an engineering process system, for example:

    trayed absorption and distillation columns, multiple adsorption beds

    operating in adsorption–desorption cycles, and membrane stages

    with interstage recompression and recycle loops. This simplified

    analysis shows clearly that the typical inherent process separation

    powers for chemical absorption technologies are much larger thanthe best alternatives currently available to treat large gas flows; 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 N2 rejection technologies the conventional cryogenic distilla-

    tion technology has a separation power more than 8 times that of N2selective adsorption and membrane processes. However, the

     Table 3

    Inherent equilibrium selectivity, ai,j, for the separation of CO2 from CH4 and N2 from CH4 by different operations, with the typical process separation power ( SP i,j) achievedin 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% CO2 to produce pipeline quality gas with 2% CO2, and (b) to reject N2 from a gas containing 4% N2  to produce a stream with 1% N2   for LNG production.

    Process Separating agent Typical inherent

    equilibrium selectivity

    Typical process

    separation power

    (a) CO2/CH4 separation   aCO2 ,CH4   SPCO2 ,CH4Amine absorption (MDEA) Liquid absorbent 860 3300

    Physical solvent (chilled methanol) Liquid absorbent 318 1900

    Adsorption—CO2 selective Solid adsorbent 2–8.5 6–22

    Membrane—CO2  selective Membrane 15–20 20–40

    (b) N2/CH4 separation   aN2 ,CH4   SPN2 ,CH4Cryogenic distillation Heat transfer 5–8 320

    Adsorption—N2  selective Solid adsorbent 1.3–2a 8–40a

    Adsorption—CH4  selective Solid adsorbent 0.25–0.5 1.05–2

    Membrane—N2  selective Membrane 2–3 2–10

    Membrane—CH4 selective Membrane 0.25–0.3 2–10

    a Inherent kinetic selectivities of narrow pore adsorbents are reported from 2 to 10, which could allow a N2 selective process with much higher separation power to be

    engineered.

    T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94–95 (2012) 123–154126

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    5/32

    separation of N2   from CH4   in adsorption-based processes is

    enhanced by the differences in rates of diffusion of N2 and CH4 that

    have been reported for small pore titanosilicate ETS-4 materials

    (Marathe et al., 2004a). Guild Associates’ Molecular GateTM pressure

    swing adsorption (PSA) process (Guild Associates, 2007) i s a

    commercial example of a N2   rejection process that relies on the

    kinetic selectivity of N2 over CH4.

    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 influence the selection of process technology for CO2/

    CH4 and N2/CH4  separations include the level of contaminants in

    the feed gas, the required level of contaminant removal or

    product purity, the flow rate and condition of the feed gas

    (temperature, pressure, water content), and for AGRUs the need

    for simultaneous or selective removal of H2S. Process selection is

    also influenced by the available disposal routes for the removed

    contaminants, which may include reinjection of CO2 for enhanced

    oil recovery (EOR) or enhanced gas recovery (EGR), and venting of 

    N2  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

    floating LNG (FLNG) plant or as a retrofit in an existing production

    plant, then the process footprint – the plan area and/or height

    occupied by the process equipment – may influence the choice of 

    process technology. Each of these process selection criteria may

    have material impacts on the feasibility, energy requirements and

    costs of CO2  and N2  removal processes.

    There are many factors that influence 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 sufficiently mature, its

    separation power to cost ratio will generally only improveasymptotically, unless a significant 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 CO2   absorption processes,

    but also introduces technologies for N2  rejection by the selective

    absorption of CH4   in hydrocarbon solvents (Mehra and Gaskin,1997) and the potential for the development of N2   selective

    solvents. Although this review is concerned primarily with CO2removal, the selection of the AGRU process is more often

    determined by the H2S removal requirements. Thus, most of the

    technical literature concerning acid gas treating focuses on the

    absorption of H2S. Commonly, the capacity of a sorbent is

    reported as an   acid gas loading capacity   which includes the

    capacity for CO2   and H2S. We have noted in   Tables 5 and 10

    which of the process technologies are suitable for the simulta-

    neous or the selective absorption of H2S.

    A large number of commercial processes are available for CO2absorption in chemical and physical solvents, including the

    technologies listed in   Table 4. Chemical absorption processes

    with amine solutions are the most commonly used acid gasremoval technologies in the natural gas industry (GPSA

    Engineering Data Book, 2004). The chemical absorption processes

    rely on reactions of the CO2   with the sorbent to form weakly

    bonded intermediate compounds, and these reactions can be

    reversed by the application of heat to release the CO2   and

    regenerate the sorbent (Olajire, 2010). Physical solvents, such as

    the mixture of polyethylene glycol–dimethyl ethers used in the

    Selexols process, selectively absorb CO2 from the natural gas feed

    according to Henry’s law so that absorption capacity increases at

    high pressure and low temperature.

    The two major cost factors in gas–liquid absorption processes

    are (1) the required sorbent circulation rate, which is determined

    by the amount of CO2  that must be removed from the feed gas

    and the CO2  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 CO2   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 CO2  capture and gas sweetening.

    Vendor/licensor Sorbent Reference

    Chemical absorption

    EconamineSM Fluor Digylcolamine, mono-ethanolamine   http://www.fluor.com

    ADIP-X Shell MDEAþaccelerator   www.shell.com

    aMDEAs 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)

    Benfield UOP Potassium carbonate   UOP Overview of Gas Processing

    Technologies and Applications, (2010)

    Catacarb Eickmeter and Associates Potassium carbonate (þorgani c addi ti ve) www.catacarb.com

    Flexsorb HP Exxon Mobil Potassium carbonate (þsteric amine)   www.exxonmobil.com

    Physical absorption

    Fluor SolventSM Fluor Dry propylene carbonate   www.fluor.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

    Sulfinol-D Shell SulfolaneþDIPAþwater   www.shell.com (Rajani, 2004)

    Amisol Lurgi Methanolþ   secondary amineþwater   Kohl and Nielsen, (1997)

    T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94–95 (2012) 123–154   127

    http://www.fluor.com/http://www.shell.com/http://www.basf.de/http://www.gasspec.com/http://www.oilandgas.dow.com/http://www.exxonmobil.com/http://www.fluor.com/http://www.uop.com/http://www.lurgi.com/http://www.lurgi.com/http://www.shell.com/http://www.shell.com/http://www.lurgi.com/http://www.lurgi.com/http://www.uop.com/http://www.fluor.com/http://www.exxonmobil.com/http://www.oilandgas.dow.com/http://www.gasspec.com/http://www.basf.de/http://www.shell.com/http://www.fluor.com/

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    6/32

    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 (NH3)

    where one or more hydrogen atoms have been substituted with

    an alkyl or aromatic group. It is the (–NH2) functional group of the

    amine molecule that provides a weak base that can react with the

    acid gases. The absorption of CO2 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 first physical absorption step is governed

    by the partial pressure of the CO2  in the gas feed. The reactions

    involved in the second step of CO2 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 CO2  absorption

    in amine treating are (Kohl and Nielsen, 1997):

    Dissociation of water:

    H2O"HþþOH (3)

    Hydrolysis and dissociation of dissolved CO2:CO2þH2O"HCO3

    þHþ (4)

    Protonation of the amine:

    RNH2þHþ"RNH3

    þ (5)

    Carbamate formation:

    RNH2þCO2"RNHCOOþHþ (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 Hþ , OH- and HCO3-

    in the amine solution affect the other reactions involving

    the amine.

    Amines can be classified according to the number of hydrogen

    atoms that have been substituted, as primary (R–NH2, where R is

    a hydrocarbon chain), secondary (R–NH–R 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 CO2   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 CO2 per

    mole of amine. However, DEA-based amine processes can achieve

    loadings of more than 0.5 mol of CO2  per mole of amine through

    the partial hydrolysis of carbamate (RNHCOO-) to bicarbonate

    (HCO3), 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 CO2  to form carbamate. Instead, CO2  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 0R 00NþCO2þH2O"RR 0R 00NHþþHCO3

    (7)

    The stoichiometry in Eq. (7) shows that theoretically tertiary

    amines can achieve a loading of 1 mol of CO2  per mole of amine,

    which is double the CO2 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 areslower than for primary and secondary amines. For some natural

    gas treating processes the slow kinetics of CO2   absorption in

    tertiary amines can be utilised to achieve selective H2S removal

    by optimisation of the contact time in the absorber to minimise

    CO2 uptake (GPSA Engineering Data Book, 2004). Alternatively, to

    enhance CO2   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 CO2(GPSA Engineering Data Book, 2004).

    Since the first application of tertiary amines in the mid-1970s,

    significant research has been directed into the further develop-

    ment of novel amine solvents. To accelerate the reaction rate of 

    tertiary amines with CO2, a primary or secondary amine can

    be included as an activator (GPSA Engineering Data Book, 2004).

    Fig. 2.  Process flow 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) 123–154128

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    7/32

    For example, the cyclic diamine piperazine has been studied as a

    promoter to improve the CO2   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 CO2  in the vapour phase, which accelerates

    the dissolution of CO2   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), whichreduces the carbamate stability. The molecular configuration

    dictates the amount of CO2   removal: severely hindered amines,

    such as ExxonMobil’s Flexsorb SE, have very low rates of CO2absorption and allows selective H2S removal. In contrast, moder-

    ately hindered amines, such as 2-amino, 2-methyl, 1-propanol

    (AMP), are characterised by high rates of CO2 absorption and high

    capacities for CO2. Weiland et al. (2010) evaluated the use of AMP

    with MEA for CO2 capture from flue gases and found it had several

    advantages including at least a 15% reduction in the required

    regeneration energy.

    A typical process flow 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 flow 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 50–70 bar (Kidnay and Parrish, 2006). The lean amine solution,

    typically an aqueous solution containing 10–65%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 theexothermic 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 0.20–0.81 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

    flash tank, to separate any dissolved hydrocarbons from the rich

    amine, and preheated to 80–105 1C before entering the stripping or

    regenerator column. In the stripping column, heat supplied by a

    steam reboiler generates vapour, which removes the CO2  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. Thevapour 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 H2S 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 CO2   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-filter 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 efficiencies

    (MacDowell et al., 2010). However, one promising strategy to

    intensify the CO2  absorption process is the use of a hollow fibre

    membrane as a gas–liquid 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 CO2   over CH4; instead the membrane provides a

    physical barrier between the gas and liquid phases, and a large

    interfacial surface area for mass transfer of CO2. 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 H2S specification

    (Kidnay and Parrish, 2006)

    yes yes yes yes

    Removes COS, CS2

    , mercaptans (Kidnay and Parrish,

    2006)

    partial partial partial partial

    50 ppm CO2  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, CO2, CS2, SO2,

    SO3, mercaptans

    some - COS, CO2,

    CS2, HCN, mercaptans

    yes - COS, CO2,CS2   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 CO2, 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) 123–154   129

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    8/32

    absorption of CO2   in the liquid solvent occurs at the liquid

    interface of the membrane. For the hollow fibre membrane to

    operate effectively as a gas–liquid contact device the pores must

    remain gas filled 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 gas–liquid membrane contactor pilot plant is reported to have

    been tested in Scotland to remove CO2 from a 5000 Nm3/h natural

    gas feed (Mansourizadeh and Ismail, 2009). Other improvementsin the performance of CO2 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 CO2-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

    (K2CO3) or sodium carbonate (NaCO3) have been employed sincethe 1950s to remove CO2   from high pressure gas streams (Kohl

    and Nielsen, 1997). The overall reactions for CO2 with potassium

    carbonate can be represented by (GPSA Engineering Data Book,

    2004):

    CO2(g)þK2CO3þH2O(l)"2KHCO3(s)   (8)

    The basic potassium carbonate process was developed by the

    US Bureau of Mines and commercialised as the Benfield 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 Benfield Process include the Catacarb Process

    (Eickmeter and Associates), which is mainly used in the ammonia

    industry, and the Flexsorb HP Process (Exxon Research andEngineering).

    The process flow diagram for a potassium carbonate absorp-

    tion system shares many features with the general amine process

    flow diagram shown in  Fig. 2. In a typical hot carbonate process

    design the absorber and stripping columns operate in a tempera-

    ture range of 100–116 1C (GPSA Engineering Data Book, 2004). If 

    H2S removal is required or if low CO2 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 CO2  removal

    to low levels for cryogenic gas processing, or GTL plant feed, the

    UOP process design can be modified 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 CO2 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 CO2 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 CO2   with

    ammonia (NH3) in solution have been developed for the capture

    of CO2 from natural gas, coal seam gas, and post combustion flue

    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 0–20 1C and (ii) processes operating with absorber at

    ambient temperatures (25–40 1C). Both variants are based on the

    same reactions of CO2  and ammonia (NH3) described by Bai and

    Yeh (1997), but at low temperatures the chilled AAP allows the

    precipitation of ammonium bicarbonate shown in Eq. (9):CO2(g)þNH3(l)þH2O(l)"NH4HCO3(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 flow scheme of an AAP plant is very similar to the

    flow 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 CO2-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

    NH3   and CO2   at temperatures greater than about 50 1C   (Dardeet al., 2010; Olajire, 2010), although temperatures of 100–150 1C

    are preferred in some designs (Gal, 2006). Typical AAP solvent

    concentrations are in the range 13–30% wt NH3 (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 CO2  capture suggests 2100–3100 kJ/

    kg CO2  compared to 3700 kJ/kg for CO2   capture by MEA,  Darde

    et al., 2010), the energy savings are not sufficiently 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 efficiency of chilled

    AAPs is only 90% (Gal, 2006), hence ammonia processes may not

    be capable of achieving very low CO2  concentrations in productgas 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 CO2.

    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 CO2  concentrations,

    the leading physical absorption technologies include the Selexols

    and Rectisols 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 flash 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 CO2sequestration, EOR or sulphur recovery.

    The Selexols process, based on a mixture of polyethylene

    glycol-dimethyl ethers, is able to remove CO2  simultaneously

    with H2S and water (GPSA Engineering Data Book, 2004). In fact,

    T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94–95 (2012) 123–154130

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    9/32

    H2

    S has a greater solubility in most organic solvents than CO2

    , a

    property that can be used to design H2S 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 CO2  for

    feed to a LNG plant, the Rectisols 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 CO2   gas (70%) is

    treated (CO2  Extraction &  Sequestration Project Riley Ridge, WY,

    2011).

    The main weakness of physical solvent technologies relative toamine 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-knownmixed-solvent processes for CO2   absorption are the Sulfinol-D

    s

    process licensed by Shell Global Solutions (Rajani, 2004) and the

    Amisol process licensed by Lurgi (Kohl and Nielsen, 1997). The

    Sulfinol-Ds process based on a mixture of Sulfolane (tetrahy-

    drothiophene dioxide), DIPA (diisopropanolamine) and water is

    capable of deep removal of CO2 to less than 50 ppm from feed gas

    containing very high concentrations of CO2   (Kohl and Nielsen,

    1997). Another variation of the Shell Global Solutions technology

    is the Sulfinol-Ms process, which is also based on Sulfolane but

    uses MDEA instead of DIPA, used mainly for selective removal of 

    H2S. The process flow scheme of the Sulfinol-Ds process is

    essentially the same as that for an amine absorption process

    (Fig. 2) with the addition of a flash 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 purification 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 CO2absorption 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 defined 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, nonflammability, 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 scientific interest in their devel-

    opment and investigation into their use in chemical engineeringapplications, 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 CO2   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 CO2  capture and natural gas sweet-

    ening has been driven by the desire to develop a solvent with a CO2capacity 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 flow diagram of a typical physical solvent process for absorption of CO2  and other acid gases from natural gas.

    T.E. Rufford et al. / Journal of Petroleum Science and Engineering 94–95 (2012) 123–154   131

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    10/32

    the most convenient ILs to synthesise, such as 1-n-alkyl-3-methy-

    limidazolium [Cnmim] cations paired with various anions including

    bis(trifluoromethylsulfonyl)amide [Tf 2N] (Bara et al., 2010b; Hughes

    et al., 2011). However, the alkyl chains (in the imidazolium cation)

    are not the ideal functional group for separating CO2 from CH4 or N2and, 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 2N] on the solubility of CO2in various ILs, with anions containing fluorinated 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 CO2 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 specific ionic

    liquids (TSIL) which incorporate an amine functional group into the

    IL, enabling it to serve as a reactive, chemical solvent for CO2. 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 CO2   per

    mole of TSIL; regeneration of the IL solvent was also achieved by

    heating the carbamate product to 80–100 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 

    CO2   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 difficult

    to synthesise and have large viscosities at ambient temperature,

    which increase further upon complexation with CO2 (Brennecke and

    Gurkan, 2010;   Karadas et al., 2010). Brennecke and co-workers

    suggest that the viscosity increase of TSILs upon complexation with

    CO2 can be eliminated through the use of aprotic heterocyclic anions

    and have filed 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 CO2 capture

    or natural gas sweetening is the use of IL þamine mixtures, in whichMEA or DEA is dissolved in an [Rmim][Tf2N] solvent; such solutions

    can have up to 116 times the CO2  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 CO2  per mole of amine and thereby reduce feed gas CO2concentrations to the ppm level, even at CO2 partial pressures below

    0.133 kPa. Such IL þamine solvents offer significantly reduced

    energy requirements relative to conventional aqueous amine sol-

    vents:  Bara et al. (2010a)  compared a IL-amine process with flash

    regeneration (similar to the process flow 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 specification of 2% CO2   for an NG feed

    containing either 15% or 5% CO2  and concluded that over a 20 yrplant life cycle, the amineþIL 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 amineþIL solvent, and the reduced duties

    required to cool, heat and regenerate the amineþIL solvent. The

    difference in the calorific properties of water and the IL, and in

    particular the duty reduction associated with not vaporizing any

    solvent during regeneration, is probably the most significant advan-

    tage of the amineþIL solvent approach to gas sweetening. A pilot

    scale unit is under construction for planned NG sweetening field

    tests in 2011–2012 (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

    solvent’s capacity, and will also degrade the reduction in regen-

    eration duty associated with an amineþIL solution. Thus in

    contrast with conventional gas processing practice, a sweetening

    process utilising an amineþIL 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 CO2. Several groups are researching the optimisation

    of these switchable solvents for improved CO2   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 CO2   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 N2  rejection processes that operate by the physical

    absorption of CH4  in a hydrocarbon oil which have been built to

    process gas feed rates of 2–30 MMscfd (AET-Technology, 2007;

    TGPE, 2009). The costs of large solvent rates and CH4   recompres-

    sion are prohibitively high for the use of CH4 selective absorption

    NRUs in large scale LNG production plants; thus there is a

    growing need for the development of a N2  selective absorbent.

    An example of a CH4  selective NRU process is that designed

    and constructed by Advanced Extraction Technologies, Inc. (AET,

    Houston USA). The flow scheme of the AET process is similar to

    the physical solvent process shown in   Fig. 3. In a typical CH4selective absorption process the feed gas is cooled before entering

    the absorption column where CH4   is dissolved into a lean oil

    solvent (TGPE, 2009). The CH4   is then recovered from the richsolvent through a series of flash vessels to produce sales gas

    containing less than 4% N2. Th e N2   from the feed gas flows

    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 CH4   absorption in the solvent.

    However, the capacity of most solvents for CH4  is relatively low

    and large solvent circulation rates are required to achieve eco-

    nomic recovery of the CH4. Furthermore, because the CH4   is

    recovered from the solvent by flashing at low pressure the gas

    must be recompressed for pipeline or cryogenic gas plant feed.

    Although most solvents have a greater solubility for CH4  than

    N2, some organo-metallic complexes (OMCs) have been reported

    to preferentially bind N2   from natural gas mixtures. Reversible

    transition metal complexes forming with N2 were first reported inthe scientific literature by Allen and Senoff (1965) who described

    the reversible reaction of hydrazine (a N2 source) with ruthenium

    chloride. Several scholarly articles describing the preparation and

    N2 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 N2 binding complexes

    to natural gas processing. The most extensive applied studies of 

    N2  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) 123–154132

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    11/32

    The absorption of N2   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 N2 molecule in an end-on configuration. 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 N2

      molecule (Bomberger

    et al., 1999). This phosphine complex absorbed up to 0.12 mmol of 

    N2/mL of solution at 1979 kPa (Alvarado et al., 1996) in equilibrium

    measurements. However, during absorption–regeneration 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 N2. The

    problems with the regeneration of the phosphine complex included

    difficulties in controlling the size of the N2-bound precipitates,

    which would bypass the regeneration system if the size was less

    than that of the phosphine solids (412mm) (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 theregeneration process are all significant.

    The patents held by Bend Research Inc. (Friesen et al., 2000,

    1993) describe OMCs based on several transition metal–ligand

    combinations including complexes with iron, as well as the perfor-

    mance of these complexes in N2 absorption–regeneration cycles. The

    patents report N2  uptakes of 0.5 mol N2/mol OMC at a N2   partial

    pressure of approximately 1130 kPa and a selectivity for N2 over CH4close 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% CO2  and 100 ppm

    O2. Although the publically available information on the Bend

    process indicates the concept of N2  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 scalegas 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 H2S. 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 CH4  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 CH4   is   94.7 1C

    (Lemmon et al., 2010), which provides a sufficient difference in

    relative volatility with N2   (NBP   195.8 1C, BP of   148.7 1C at3150 kPa) for separation of CH4þN2   mixtures by cryogenic

    distillation. Relevant to such separations are vapour–liquid equi-

    librium curves, such as the one shown in   Fig. 5   for N2–CH4   at

    conditions accessible in a cryogenic LNG plant. The equilibrium

    curve in Fig. 5 represents a typical relative volatility aN2,CH4  valueof 5–8. To illustrate the separation of N2 and CH4 in the cryogenic

    distillation column, we have applied the McCabe–Thiele method

    (McCabe and Thiele, 1925) to estimate the number of ideal

    equilibrium stages required to produce a liquid CH4   bottoms

    product of 95% from a feed containing 50% N2 and 50% CH4. These

    feed and production compositions were selected here to allow

    illustration of the separation process; in a real process to produce

    a CH4   containing less than 1% N2   and with a high rate of CH4

    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). R ¼generic organic functionality.

    Fig. 5.   Illustration of the binary CH4–N2   vapour–liquid equilibrium relationship

    and the construction of a McCabe–Thiele diagram to calculate the number of ideal

    equilibrium stages for separation by distillation. The vapour–liquid equilibrium

    data shown here represents a relative volatility,   aN2CH4 , of 7 (calculated using

    REFPROP (Lemmon et al., 2010) for an operating pressure of 2757 kPa. In this

    hypothetical example, the feed contains 50% N2 with product specifications of 5%

    N2 in the CH4-rich bottom product and 5% CH4 in the N2-rich overheads product.

    These compositions were selected as an example which could be shown clearly on

    this figure and do not represent a typical set of operating conditions. To meet more

    realistic processing objectives, such as a higher purity CH4-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) 123–154   133

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    12/32

    distillation column would be higher than the eight ideal trays

    shown in the McCabe–Thiele construction on  Fig. 5. (The corre-

    sponding graph would be more difficult to read.) In practice,

    modern cryogenic NRUs (with   SP N2,CH4  320) can produce very

    high purity CH4 at high recovery rates, which also reduces the CH4content in the N2   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 N2  rejection methodthat has been demonstrated at gas flows above 25 MMscfd to

    achieve very high methane recovery (typically above 98%) and high

    purity N2   (approximately 1% CH4). The selection, and optimised

    design, of a cryogenic NRU is principally determined by the

    concentration of N2 in the feed gas (Wilkinson and Johnson, 2010).

    The feed gas pressure, feed gas flow rate, concentration of con-

    taminants and product specifications (sales gas or LNG) also

    influence process selections. The optimum design of cryogenic

    NRUs, like for most cryogenic processes, is an exercise in balancing

    the energy efficiency and process flow sheet integration to reduce

    the power consumption required for compression of the CH4refrigerant 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, CO2 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 N2   vapour (typicallyo1% CH4) is drawn from the column overheads and the CH4-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 CH4  heat-pump cycle driven by an external com-

    pressor provides the reboiler and condenser duties, with the

    closed-loop CH4   condensed at a high pressure in the reboiler

    and revaporised at low pressure in the condenser.

    As the N2 content in the feed gas increases, the CH4 in the upper

    stages of the column becomes more difficult to condense. The

    operating flexibility of a single-column NRU process is limited by

    (1) the critical pressures of nitrogen–methane mixtures, which

    limits the maximum pressure of the distillation column to approxi-

    mately 2800 kPa, and (2) the minimum practical temperature of CH4

    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% N2.

    A double-column NRU can provide additional process flexibility

    compared to the single-column process to allow the separation of 

    gases containing higher N2 concentrations or gases in which the feed

    gas quality varies. In the double-column N2   rejection process the

    NRU feed gas is cooled, throttled and fed to a high pressure (HP)

    column operating typically at 1000–2500 kPa (Agrawal et al., 2003).

    Having had some of the N2   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

    N2-rich vapour from the HP column is condensed to provide refluxfor both the high pressure column and the low pressure column. The

    low pressure column produces a high purity N2 stream (o1% CH4);

    which is used to cool N2   reflux 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 CH4-rich bottom product of the LP column is

    pumped through the crude sub-cooler, which provides the LP

    column’s reboiler duty by condensing the overheads of the HP

    column, vaporised and reheated against the feed gas. Alternatively,

    for LNG production the CH4-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 N2concentration of the feed gas. Importantly, the prefractionation

    column reduces the volume of N2-rich gas that must be processed

    at low temperatures in the main NRU column(s) and this reduc-

    tion in gas volume can significantly 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

    liquid–vapour separator and a low pressure column.

     3.2. Low-temperature CO 2 removal processes

    The separation of CO2  from natural gas by low-temperature

    processes (operating at temperatures below 0 1C) can be

    Fig. 6.   A process flow schematic of a typical single-column N2   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) 123–154134

  • 8/18/2019 The Removal of CO2 and N2 From Natural Gas Review (Rufford Et Al 2012)

    13/32

    categorised as: (1) gas–liquid phase separations operating at

    temperatures above the CO2   triple point temperature of 

    56.6 1C (Lemmon et al., 2010) and (2) gas–solid phase separa-

    tions where desublimation of CO2  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 CO2capture technologies, most of the processes operate at tempera-

    tures above the scientific definition of cryogenic as   153 1C

    (Agrawal et al., 2003;   Radebaugh, 2007). To overcome theproblems associated with the formation of CO2   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 ZoneTM and CryoCells processes).

    Other low-temperature CO2   removal technologies under devel-

    opment include systems in which mechanical methods are used

    to separate the CO2 rich phase from the natural gas. For example,

    Willems et al. (2010)   report the  C3sep  (condensed contaminant

    centrifugal separation) process in which condensed CO2  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 CO2  on mechanical fins. We focus our discussions in this

    review on the commercialised Ryan/Homes process and the pilot-

    plant demonstrated Controlled Freeze ZoneTM (along with the

    similar CryoCells process).

    The extractive distillation approach to solving the problem of 

    CO2 freezing in CH4–CO2 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 C2–C5 alkane) to the condenser of the distilla-

    tion column shifts the operation away from conditions that favour

    solids formation, because the solubility of CO2 in the liquid phase

    can be increased, the overheads temperature can be raised, and

    the column can be operated at a higher pressure since themixture’s critical pressure increases. A typical four column

    Ryan/Holmes process configuration incorporates a de-ethaniser

    column, a CO2  recovery column, a demethaniser column, and a

    column for recovery of the hydrocarbon additive. Further details

    on Ryan/Holmes configurations and operating issues for separa-

    tions of methane–CO2, ethane–CO2, and CO2–H2S are discussed in

    the  GPA Engineering Data Book.

    The Controlled Freeze ZoneTM (CFZTM) process was first patented

    by ExxonMobil in 1985 (Valencia and Denton, 1983) and tested in a

    Texas pilot plant during 1986–1987 (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 ZoneTM—increasing the supply of clean burning

    natural gas, 2010). The separation tower of the CFZTM process is splitinto three sections with an upper rectification section and a lower

    stripping section (both conventional distillation sections) separated

    by the CFZTM section, as shown in Fig. 7 (Fieler et al., 2008). In the

    CFZTM section, the liquid falling from the rectification section is

    contacted with a cold methane stream (90 to   85 1C), which

    causes the CO2 to freeze out of the methane mixture. The CO2 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 CFZTM process can

    produce pipeline quality gas, and when implemented with a

    modified rectification section is claimed to be capable of producing

    a sweet gas of less than 50 ppm CO2  (Nichols et al., 2009).

    Cool Energy’s CryoCells 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 CFZTM process, the CryoCells process

    operates by the controlled freezing and subsequent remelting of 

    CO2. The basic thermodynamic path for the CryoCells operation

    involves cooling a dry, feed gas (at 5600–6600 kPa) to just above

    the CO2 freezing point (for example to  60 1C) to condense some

    or all of the vapour, followed by an isenthalpic flash to furthercool the mixture to obtain solids, liquid CO2   and a CH4-rich

    vapour. Pilot plant trials of the CryoCells process demonstrated

    the production of