flue gas treatment via co2 adsorption.pdf

Upload: gaye

Post on 01-Jun-2018

232 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    1/15

    Chemical Engineering Journal 171 (2011) 760–774

    Contents lists available at ScienceDirect

    Chemical Engineering Journal

     journa l homepage: www.elsevier .com/ locate /ce j

    Flue gas treatment via CO2 adsorption

    Abdelhamid Sayari a,b,∗, Youssef Belmabkhouta, Rodrigo Serna-Guerrero b

    a Department of Chemistry, University of Ottawa, 10Marie Curie, Ottawa, ON,Canada K1N6N5b Department of Chemical andBiological Engineering, University of Ottawa, 161 Louis Pasteur, Ottawa, ON,Canada K1N6N5

    a r t i c l e i n f o

     Article history:

    Received 20 June 2010

    Received in revised form 14 January 2011

    Accepted 5 February 2011

    Keywords:

    CO2  adsorption

    Zeolite

    Carbon

    MOFs

    Supported amines

    a b s t r a c t

    Adsorption separation has gained considerable attention as a viable alternative to the currently used,

    high energy-demanding aqueous amine scrubbing technologies. This review is a summary of  the main

    contributions regarding the development of new adsorbents for post-combustion CO2 capture. Emphasishas been placed on materials evaluated at representative flue gas conditions of CO2 partial pressure (i.e.,

    0.05–0.2 bar) and temperature (25–75 ◦C). Whenever possible, the effect of  moisture on the adsorbent

    stability and CO2 uptake is included, although relatively few studies in the literature have focused on this

    issue. This review includes adsorbents produced by modification of existing commercial materials as well

    as newly developed materials. These adsorbents were separated in two major classes, namely (i) physical

    adsorbents including carbons, zeolites and metal-organic frameworks and (ii) chemical adsorbents, i.e.,

    amine-functionalized materials. A critical analysis of  the literature is provided with the aim of tracing

    the main paths currently pursued toward the development of suitable CO2 adsorbents and to provide a

    general overview of the advantages and limitations of each family of adsorbents.

    © 2011 Elsevier B.V. All rights reserved.

    1. Introduction

    As the concentration of carbondioxide (CO2) in the atmospherekeeps increasing, serious concerns have been raised with respect

    to its impact on the environment. Since it started being monitored

    in 1958, the increase of CO2  concentration in the atmosphere has

    accelerated from less than 1ppm/yr prior to 1970 to more than

    2 ppm/yr in recent years [1]. As a result, the atmospheric level of 

    CO2   increased from 315ppm in 1958 to 385ppm in 2009 [1,2].

    CO2   is considered to be the main anthropogenic contributor to

    the greenhouse gas effect, as it is allegedly responsible for 60% of 

    the increase in atmospheric temperature, commonly referred to as

    global warming [2,3]. Among the various sources of CO2, approxi-

    mately 30% is generated by fossil fuel power plants, making them

    major contributors to global warming [4]. Despite their impact on

    theenvironment, it is acknowledgedthat fossil fuelswill remain the

    leading source of energy for years to come, for both power gener-ation and vehicle transportation. Therefore, it is critical to develop

    effective methods for the capture and sequestration of CO2   from

    post-combustion effluents, such as flue gas. Some reviews dealing

    with the main sources of CO2  and potential strategies to prevent

    theirrelease to theenvironment areavailable in the literature [5,6].

    Gasabsorption using alkanolaminesolutions has been used forCO2

    ∗ Corresponding author at: Department of Chemistry, University of Ottawa,

    10 Marie Curie, Ottawa,ON, Canada K1N 6N5.

    E-mail address: [email protected] (A. Sayari).

    scrubbing on industrial scalefor decades. However,this process has

    a number of shortcomings. For example, it generates severe cor-

    rosion of the equipment, and the regeneration of amine solutionsis highly energy intensive [7]. These drawbacks have been widely

    documented, prompting a search for alternative technologies. One

    viable route is adsorptionwhich, comparedto other separationpro-

    cesses, is recognized to be attractive to complement or replace the

    current absorption technology due to its low energy requirement

    [4,8]. Therefore, the use of appropriate adsorbents may potentially

    reduce thecost associatedwith CO2 separationin theoverallcarbon

    capture and storage (CCS) strategy.

    Suitable adsorbents for CO2 removal from flue gas should com-

    bine several attributes, including:

    (i) High CO2   adsorption capacity: CO2   equilibrium adsorp-

    tion capacity is one of the main properties used to screen

    new adsorbents. Knowledge of the equilibrium adsorption

    isothermsisofprimeimportanceforearlyevaluationofpoten-

    tialadsorbents.Whenever possible,this review willbe focused

    on adsorption properties measured under conditions relevant

    to flue gas treatment, i.e., less than 0.4 bar CO2  partial pres-

    sure with a total gas pressure of 1–2bar and temperature

    below 70–80 ◦C.As a ruleof thumb,Hoet al. [9] suggested that

    an optimum adsorbent for CO2 capture from flue gas, should

    exhibit a CO2  adsorption capacity of 2–4mmol/g. It is well

    established that from the slope of the adsorption isotherm at

    low pressure, it is possible to estimate the adsorbate affin-

    ity for a given adsorbent. Thus, in terms of CO2   uptake, the

    1385-8947/$ – seefront matter © 2011 Elsevier B.V. All rights reserved.

    doi:10.1016/j.cej.2011.02.007

    http://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.cej.2011.02.007http://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.cej.2011.02.007http://www.sciencedirect.com/science/journal/13858947http://www.elsevier.com/locate/cejmailto:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.cej.2011.02.007http://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.cej.2011.02.007mailto:[email protected]://www.elsevier.com/locate/cejhttp://www.sciencedirect.com/science/journal/13858947http://localhost/var/www/apps/conversion/tmp/scratch_3/dx.doi.org/10.1016/j.cej.2011.02.007

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    2/15

     A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760–774 761

    ideal materials shouldexhibit a CO2 adsorption isotherm with

    steep slope (favorable CO2  adsorption isotherm) correspond-

    ing to high uptake at low CO2   partial pressure. A less steep

    slope (unfavorable CO2 adsorption isotherm) is indicative of a

    lower affinity toward CO2.

    (ii) Fast kinetics: Adsorption kinetics affects primarily the work-

    ing adsorption capacity in dynamic processes such as

    adsorption in a fixed bed column. A suitable CO2  adsorbent

    will have a high rate of adsorption, resulting in a working

    capacity close to equilibrium capacity over a wide range of 

    operatingconditions.However, determinationof kinetic prop-

    erties such as diffusionis one of themost challengingissues in

    adsorption science, as it involves parameters not always read-

    ily available, such as particle size of the adsorbent and use of 

    adequate experimental set-ups and conditions.

    (iii) High CO2 selectivity: The adsorbentselectivitytoward CO2 has

    a direct impact on the degree of purity of the product. This in

    turn, plays a major role in the economics of the CO2 adsorption

    process [9]. Ideally,anadsorbentforfluegastreatmentwillnot

    adsorb any nitrogen.

    (iv) Mild conditions for regeneration: The ease of regeneration of 

    the adsorbent is a key property in the selection of materials

    for CO2   separation. Depending on the structural and chemi-

    calproperties of the adsorbent, adsorption–desorption cyclingmay be achieved via temperature, pressure (or vacuum), con-

    centration swing adsorption or a combination thereof. In

    practice, incorporation of functional groups can be used to

    modify the adsorbent–adsorbate interactions (e.g., Van der

    Waals, electrostatic, hydrogen bonding or acid–base interac-

    tions) and affect the CO2   uptake and selectivity. Optimum

    interactions should be neither too weak nor too strong. Too

    weak bonding results in low CO2   adsorption capacity at low

    pressure, but easy regeneration. Conversely, strong bonding

    induces high adsorption capacity but desorption will be diffi-

    cult and costly.

    (v) Stabilityduring extensive adsorption–desorption cycling: The

    lifetime of adsorbents, which determines the frequency of 

    theirreplacement, is a critical property of equal importance asthe CO2 adsorption capacity, selectivity and kinetics, because

    of its direct impact on the economics of any commercial scale

    operation.

    (vi) Tolerance to the presence of moisture and other impurities in

    the feed: In addition to CO2   and N2, flue gas contains water

    vapor and other impurities such as O2   and SO2. The degree

    of tolerance and the affinity of the adsorbent to such impu-

    rities may affect significantly the strategy to be used, with

    direct impact on the overall economics of the CO2   separa-

    tion process. Moisture is knownto adverselyaffect CO2 uptake

    in a variety of physical adsorbents such as zeolites and acti-

    vated carbon. Consequently, the strategies proposed for CO2adsorption from flue gas is likely to include an upstream dry-

    ing step. As a result, the overwhelming majority of publishedreports dealing with physical adsorbents have not examined

    moisture effects. Whenever possible, this review will provide

    a general picture on the behavior of the adsorbents in the

    presence of moisture. It is also generally established that CO2adsorbents have high affinity to SO2   and even some affinity

    toward NOX, which may adversely affect the CO2   adsorp-

    tion capability of the material. Thus, abatement of SO2   and

    NOX   from flue gas prior to CO2   capture is required in most

    cases.

    (vii) Low cost: This is another important parameter to be con-

    sidered in the development of any potential adsorbent. At

    this stage, information on adsorbent cost and other economic

    considerations are rather scarce in the literature. Thus, cost-

    related issues will not be discussed in this review.

    Because flue gas is generally cooled down to ca. 55◦C, to allow

    appropriate conditions for SO2 and NOX abatement [10,11], when-

    ever possible, this review will be focused on literature reports

    dealing with CO2  adsorption using 5–20% CO2-containing mixtures

    with a total pressure of 1–2 bar, and temperature between 25 and

    70 ◦C. Fora more comprehensiveaccounton CO2 adsorbents in gen-

    eral, the reader may refer to an excellent review by Choi et al. [12].

    Similarly, the field of high temperature CO2 capture has also been

    reviewed by Lee et al. [13].

    Adsorbents for CO2  capture can be categorized in many ways,

    based on their chemical composition, structural characteristics or

    according to the adsorption mechanism involved, i.e., physical vs.

    chemical.Physical adsorbents for CO2 capture include carbon mate-

    rials, alumino-silicas such as zeolites, alumino-phosphates (AlPOs)

    and alumino-silico-phosphates (SAPOs), and more recently metal

    organic frameworks (MOFs). The CO2   chemical adsorbents dis-

    cussed in this review refer to those obtained through incorporation

    of amine groups into solid supports such as mesoporous silica.

    Consequently, in this review we distinguished two classes of CO2adsorbents for stack gas treatment, namely physical and amine-

    functionalized adsorbents.

    2. Physical adsorbents

     2.1. Carbons

    Because of their wide availability, low cost and high ther-

    mal stability, it is largely established that activated carbons have

    advantages over other CO2   adsorbents. Among the carbon based

    adsorbents reported in the literature, activated carbons (ACs) and

    carbon nanotubes (CNTs) are the most investigated materials. CO2adsorption on activated carbons has been studied experimentally

    and theoretically for a long time [14] and has found commercial

    applications[15,16]. There is a wide rangeof activated carbons with

    a large variety of microporous and mesoporous structures. Acti-

    vated carbon may be produced from many raw materials such as

    coal, coke pitch, wood or biomass sources (e.g., saw dust, coconutshells, olive stones), often via two steps: carbonization and acti-

    vation [17]. Carbon molecular sieves (CMS), which are a sub-class

    of activated carbon with narrow pore size distribution (PSD), are

    kinetic-based adsorbents. They have been commercialized mainly

    for the separation of air and the production of high purity N2[18,19]. However, at low CO2  partial pressure, activated carbons

    exhibit lower adsorption capacity and selectivity than zeolites due

    mainly to their less favorable adsorption isotherms. In spite of 

    the hydrophobic character of carbon-based adsorbents, their CO2adsorption ability is adversely affected by the presence of water

    vapor [20].

    Table 1 shows literature data on CO2   adsorption capacity and

    selectivity of activated carbons and carbon nanotubes in the par-

    tial pressure range of 0.1–0.4 bar at 298–333 K. Considering 1 and2 bar as thelowest andhighesttotalpressureof fluegas, the0.1 and

    0.4 bar were chosen arbitrarily as the lowest and the highest CO2partial pressure relevant to flue gas treatment. Notice that most

    studies dealing with CO2   adsorption on activated carbons were

    undertaken at high pressure and room temperature.

    It is important to notice that, although adsorption capacity

    varies considerably for different activated carbons at high pressure

    [24,25], the adsorption capacity at low pressure is less sensitive to

    the nature of carbons. As seen in Table 1, the typical CO2  equilib-

    rium adsorption capacityfor activatedcarbons at a partial pressure

    of 0.1bar is 1.1mmol/g at room temperature but decreases rapidly

    to 0.25mmol/g at 328K. In terms of CO2 adsorption capacity, acti-

    vated carbons may be particularly interesting for CO2 removal but

    only at high pressure. For example, Himeno et al. [24] showed

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    3/15

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    4/15

     A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760–774 763

     Table 2

    Literature survey on CO2 adsorption properties of some zeolites and zeolite-like materials at low pressure.

    Zeolites/Si/Al ratio CO2 adsorption

    temperature (K)

    Adsorption capacity at

    0.1–0.4 bar (mmol/g)

    N2  adsorption capacity

    at 0.9–1.6 bar (mmol/g)

    CO2 /N2  capacity molar

    ratio

    Reference

    NaX/1 298 2.8–3.9 0.264–0.46 11–8.5 [48]

    NaX/1 323 1.43–2.49 – – [48]

    LiX/1 303 3.1–4.6 – – [44]

    NaY/2.4 323 0.45–1.17 – – [49]

    CsY/2.4 333 0.86–1.2 – – [46]

    KY/2.4 333 0.75–1.6 – – [46]Silicalite/∞   334 0.16–0.45 0.1 1.6 [50]

    H-ZSM-5/30 313 0.7–1.5 0.23 3 [51]

    Li-MCM-22/15 333 0.68–1 – – [52]

    –, notavailable.

    favorable adsorption isotherm. This was explained by the domi-

    nantacid–base (CO2-framework oxygen atom) interaction over the

    polarizing effect in the case of CsY and KY faujasites (particularly

    for CsY and to a lesser extent for KY), in contrast to LiY, NaY and X

    faujasites.

    Table 2 shows the CO2   adsorption properties of different zeo-

    lites and zeolite-like materials. As seen, the adsorption capacity

    decreased drastically when the temperature increased from 298

    to 323K. Akten et al. [47] showed that the CO2

    /N2

      selectivity for

    Na-4A type zeolite also decreased at increased temperature.

    In terms of CO2  adsorption kinetics, zeolites are ranked among

    the fastest adsorbents, reaching equilibrium capacity within min-

    utes [12]. Moreover, a large number of studies were devoted to

    NaX faujasite using different recycling configurations, including

    temperature swing and pressure swing adsorption [9,40,41,53].

    Although the CO2   adsorption enthalpy on X and Y zeolites was

    found to be dependent on the nature of extraframework cations,

    within the range of 30–50 kJ/mol, it is low enough to allow

    reversible CO2  adsorption. Zeolites generally operate without any

    loss in performance, provided that the feed stream is strictly dry.

    Although low silica materials exhibit high adsorption capacity and

    selectivity at low pressure with favorable isotherms, they are very

    sensitive to the presence of water, which strongly inhibits the

    adsorption of CO2  [54]. This prompted some investigations on theability of hydrophobic high silica zeolites such as MWW zeotype

    [52] and NaZSM-5 [55] to remove CO2. However, because high

    silica microporous materials contain less extraframework cations

    than faujasite zeolites, they exhibit lower adsorption capacity and

    adsorption enthalpy [45]. Moreover, similarly to X and Y zeolites,

    theyshow decreasing selectivity at increasing temperature [47,56].

    Fig. 2. Adsorption isotherms of CO2

      on cation-exchanged Y faujasites [46].

    Alumino-phosphates (AlPO) and silica-alumino-phosphates

    (SAPO) are another class of zeolitic materials that were investi-

    gated as potential CO2  adsorbents [57,58]. The overall framework

    of AlPOs is neutral and is expected to behave as silicalite or dea-

    luminated Y faujasite for CO2 as was shown by Deroche et al. [58]

    usinga combinationof molecularsimulation andmicrocalorimetry.

    In fact AlPO-18 (with AEI structure) exhibits unfavorable adsorp-

    tion isotherm. Similarly to NaX and NaY, the framework of some

    SAPOs is negatively charged and the overall charge is balanced by

    extraframework cations. In this case, it is expected to obtaina more

    favorable CO2 adsorption isotherm with higher adsorption capac-

    ity at low pressure of CO2  as reported by Castro et al. [57] f or the

    proton form of SAPO-34. However, the CO2 adsorption capacity on

    SAPO remains lower than X and Y faujasites.

    In conclusion, because of their often highly favorable CO2adsorption isotherms, zeolites and zeolite-like materials with low

    Si/Al ratios are among the most promising adsorbents for CO2 cap-

    ture from flue gas. However, because of their highly hydrophilic

    character, the flue gas needs extensive drying prior to CO2 capture.

    Notice that among zeolites, 13X is has been the most investigated

    material for the purpose of CO2 capture [9,40,41,53,59]. As pointed

    out by Ho et al. [9], further work to develop more selective zeolite

    adsorbents toward CO2 vs. N2 and O2 may reduce considerably the

    cost of CO2 capture.

     2.3. MOFs and zeolite-like MOFs

    Although an emerging class of porous materials, metal organic

    frameworks (MOFs) have attracted a growing interest, motivating

    extensive studies on their CO2 adsorption properties, both theoret-

    ically and experimentally. MOFs are porous crystalline materials

    composed of self-assembled metallic species and organic linkers

    [60,61]. Their pore size and shape can be easily tuned by chang-

    ing either the organic ligands or the metallic clusters. They are

    typically rigid materials, but some of them exhibit structural flex-

    ibility upon adsorption and desorption of gases or liquids [62,63].

    The wide range of MOFs combined with their desirable proper-ties such as their remarkably high surface area and controlled

    pore size and shape, prompted extensive work on their adsorp-

    tive properties, particularly for storage of light gases (H2, CH4) and

    storage and separation of CO2. Although the majority of inves-

    tigations on CO2   adsorption over MOFs used pure CO2, as well

    as CO2-containing mixtures, most measurements and simulations

    were carried out at high pressure and often at room or subambient

    temperature. Seminal contributions in the synthesis of novel MOFs

    andtheir CO2 adsorptionproperties werereportedby Millward and

    Yaghi [64]. Their early work was followed by an extensive effort to

    develop new types of MOFs for the separation and storage of CO2[65–72]. Millward and Yaghi [64] showed that MOF-117 exhibits

    an unprecedented CO2  adsorption capacity at high pressure (e.g.,

    ca. 150 wt% at 40 bar), but very small CO2   uptake at subatmo-

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    5/15

    764   A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760–774

     Table 3

    Literature survey on CO2 adsorption properties of some MOFs and ZMOFs.

    MOFS   Temperature (K) CO2  adsorption capacity

    (mmol/g at 0.1–0.4 bar)

    N2  adsorption capacity

    (mmol/g at 0.9–1.6 bar)

    CO2/N2  selectivity Reference

    MOF-508 323 0.1–0.7 0.6–0.9 2 [75]

    Cu-BTC 298 0.5–2 0.25 15 [79]

    MIL-53 303 0.5–1.15 – – [67]

    Ni/DOBDC 296 2.7–4.01 – – [72,73]

    CO/BOBDC 296 2.8–5.36 – – [72,73]

    Mg/DOBDC (Mg-MOF-74) 296 5.36–6.8 – – [68,72]ZIF-78 298 0.77–1.36 – 50   [70,80]

    –, notavailable.

    spheric pressure. More recently, Caskey et al. [72] reported a much

    higher and reversible adsorption capacity for pure CO2  (23.6wt%,

    5.36mmol g−1) at 0.1 bar and room temperature on Mg/DOBDC (or

    Mg-MOF-74) (Fig. 3). Although this is an excellent finding, adsorp-

    tion of CO2  in mixtures with N2  was not reported. Additional data

    regarding adsorption of CO2  at 0.1bar on a number of MOFs may

    be found elsewhere [73].

    It is important to mention that the majority of MOFs exhibit

    unfavorable adsorption isotherms for CO2   in the low pressure

    range. Moreover most of these materials adsorb considerable

    amounts of N2, leading to low selectivity toward CO2. The highestselectivity toward CO2 vs. N2 was in the range of 5–30 [74–76], and

    generally the CO2   adsorption capacity is dramatically reduced at

    higher temperature, accompanied by a drop in the CO2 adsorption

    selectivity. In terms of kinetics, MOFs are as fast CO2 adsorbents as

    zeolites according to somecomputational studies [66–77]. Basedon

    the aforementioned observations, MOFs seem to be more suitable

    for CO2 storage rather than separation.

    Table 3 shows the adsorption properties of different types of 

    MOFs at low pressure. As seen, the adsorption capacity as well as

    the CO2/N2 selectivity formost MOFs, were very lowand decreased

    drasticallywhen the temperature increasedfrom 298to 323K. This

    has been documented byBarcia etal. [78], Bastinetal.[75], Baeetal.

    [74], Yang et al. [79] who showed decreasing CO2/N2 selectivity at

    increased temperature for MOF-508b and Cu-BTC.MOFs and ZMOFs structural, chemical and thermal stability has

    beenhardlyaddressedin the literature, until recently.It is generally

    recognizedthatby farthemostcritical issuefor the stabilityof these

    materials is their hydrothermal stability. The behavior of MOFs

    and their subfamilies in hydrated conditions varies widely, from

    materials that irreversibly degrade even under mild conditions to

    materialsthatare highlystable in boiling water. Forexample,MOF-

    Fig. 3. CO2 adsorption isotherms(296K, 0–1atm)for M/DOBDC materials. Inset is a

    close-up of the low pressure region. Filled and open symbols represent adsorption

    and desorption data, respectively [72].

    117 and IRMOF-1 were reported to be unstable upon exposure to

    air in the presence of humidity [81–83]. The concerns raised by the

    stability of MOFs prompted the discovery of a new class of MOFs

    referred to as zeolite-like MOFs (ZMOFs) or zeolitic imidazolate

    frameworks (ZIF). ZMOFs are crystalline porous materials which

    combine the highly desirable properties of zeolites and MOFs,

    such as microporosity, high surface areas, and exceptional ther-

    mal and chemical stability [69–81]. Because of the strong bonding

    between the imidazolate linker and the metal center, many ZMOFs

    have high thermal (>673 K) and moisture resistance compared to

    other MOF structures [84,85]. Although significant improvementwas observed in terms of CO2  v s. N2  adsorption selectivity, which

    increased up to 50 for ZIF-78 [80], theCO2 adsorption capacity was

    still low at low CO2 partial pressure (Table 3).

    Zeolites, MOFs and ZMOFs are typically hydrophilic and their

    application for CO2   capture from flue gas requires partial or com-

    plete drying of the gas stream. To circumvent this limitation,

    new materials with no hydrophilic adsorption sites referred to

    as covalent organic frameworks (COFs) were developed. COFs are

    crystalline organic porous materials without metal ions. Furukawa

    and Yaghi [86] and Babarao and Jiang [87] reported high CO2adsorption capacities for a series of COFs, but adsorption at low

    partial pressure of CO2 appeared to be significantly lower than for

    Mg-MOF-74.

    In summary, MOFs, ZMOFs and COFs may be promising materi-als for CO2  removal provided that more favorable CO2  adsorption

    isotherms are obtained. Their selectivity and capacity at low par-

    tial pressure of CO2  in gas mixtures are quite low and more likely

    to be suitable for CO2 storage rather than CO2 separation from flue

    gas. Although in their early stages of development, MOFs, ZMOFs

    and COFs are promisingmaterials for CO2 adsorption showing very

    interesting and adjustable properties.

    3. Amine-functionalized adsorbents

    The technology currently used in industry for CO2  capture is

    absorption with liquid amine solutions. The removal of CO2   by

    amines occurs viathe widely accepted formation of carbamate and

    bicarbonate species, as represented in Scheme 1 [88]. These are

    reversible reactions that permit the regeneration of amines, typi-

    cally by heating the CO2-rich solution.

    2(RNH2)+ CO2  ↔ RNHCO2−RNH3

    +

    carbamate

    RNH2 + CO2 +H2O ↔ RNH3+HCO3

    bicarbonate

    RNH2←→(RNH3

    +)2CO32−

    carbonate

    The liquid amine absorption process inspired researchers to use

    amine-modified solid materials as adsorbents for CO2  capture. As

    far as flue gas treatment is concerned, it was anticipated that sup-

    ported amines will maintain a high selectivity toward CO2  with a

    negligible uptake of other components, particularly N2, but with-

    out the aforementioneddrawbacks associatedwith aqueous amine

    solutions. According to a study on amine-functional adsorbents by

    Gray et al. [89], a capacity of at least 3 mmol/g is required for this

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    6/15

     A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760–774 765

    Scheme 1. Typical reaction pathway between CO2   and amines [88].

    process to be competitive against absorption technologies, corrob-

    orating the previously mentioned range of 2–4 mmol/g proposed

    by Ho et al. [9]. Although the early efforts to produce amine-

    functionalized adsorbents were not particularly successful in terms

    of adsorption capacity, the collective effort of several research

    groups resulted in significant performance improvements, leading

    to increasing interest in this subject matter. Based on ISI Web of 

    Knowledge, Fig. 4 illustrates the remarkable increase in the number

    of publications related to CO2 adsorption on amine-functionalized

    materials, with ca. 70% of these contributions published in the last

    6years.

    We have broadly organized the present section according to

    the type of interactions between amine groups and the support,

    namely (i) amine-impregnatedmaterials where mostly weakinter-

    actions occur, and (ii)covalently bonded amine-containing species,obtained typically via surface-grafting of aminosilanes. The ratio-

    nale behind such classification is that materials with either strong

    or weak interactions exhibit a number of common characteristics.

    An example is that grafted materials offer comparatively higher

    rate of adsorption than amine-impregnated adsorbents [91] and,

    in some cases even higher than commercial adsorbents such as

    13X [90]. However, the organic content of amine-grafted adsor-

    bents depends on the surface density of hydroxyl groups, needed

    to anchor the aminosilane. As for impregnated amines, higher load-

    ingsmay be achieved, butoften accompaniedby increasinglystrong

    diffusion limitations.

     3.1. Amine-impregnatedmaterials

     3.1.1. Ordered mesoporous supports

    Xu et al. [92] were first to report on polyethyleneimine (PEI)-

    impregnated mesoporous materials for CO2   adsorption, coining

    the term “molecular basket”. It was found that the adsorption

    capacity of PEI-impregnated MCM-41 improved at increased load-

    ing. The highest value of adsorption capacity, corresponding to

    3.02mmol/g was obtained under a stream of pure CO2   at 75◦C

    using a sample with 75w t% PEI. However, the maximum effi-

    ciency, i.e., CO2/PEI molar ratio was obtained in the presence of 

    Fig.4. Number of publications relatedto CO2 capture onamine-functional materials

    according to ISI-Web of Science database.

    a material containing 50wt% PEI, and decreased steadily at higher

    loadings. As discussed later, this behavior was confirmed by other

    researchers. Further work under conditions relevant to flue gas

    treatment showed that the 50% PEI on MCM-41 silica exhibits an

    adsorption capacityof ca.2.1mmol/g inthe presence of 10% CO2/N2at 75 ◦C. A particularly interesting behavior of PEI-impregnated

    MCM-41 materials was the fact that, unlike other adsorbents,

    adsorptioncapacityimproved as temperature increased from 25 to

    75 ◦C. Since theactual adsorption event is exothermic in nature, the

    increasing adsorption capacity with temperature was attributed

    to the occurrence of a bulk-like state of PEI inside the meso-

    pores, withaminegroups notreadilyaccessible at lowtemperature,

    resulting in a diffusion-limited process. Since then, other authors

    working on PEI-impregnated materials reported similar findings

    and the idea of a diffusion-limited process has been generallyaccepted.

    Following their early study on PEI-impregnated MCM-41, Xu

    et al. [93] analyzed theadsorption of CO2 inhumidstreams. A posi-

    tive effectof moisture was observed in terms of increased capacity,

    particularly when the molar concentration of water was equal or

    lower to that of CO2, providing support to the bicarbonate forma-

    tion mechanism. No further increase in adsorption capacity was

    observed for streams with higher moisture content. Accordingly,

    itscapacitywas enhanced from 2.01 mmol/g in a simulateddry flue

    gas containing 15% CO2 to 2.84 mmol/g in a stream containing 10%

    moisture and13% CO2, balance air. In a later contribution, thesame

    group [94] impregnated PEI on SBA-15 under the assumption that

    the structural characteristics of the support would affect the per-

    formance of the aminated adsorbents. Allegedly, due to the largerpore size and volume of SBA-15 compared to MCM-41, PEI-SBA-15

    used the amine groups more efficiently under the same loading of 

    50wt% PEI. Indeed, as mentionedabove, while PEI-MCM-41 exhib-

    ited an adsorption capacity of 2.1 mmol/g, PEI-SBA-15 showed a

    capacity of 3.18 mmol g−1 under a flow of 15% CO2   balance air at

    75 ◦C.

    Ahn’s group [95,96] has also investigated the adsorptive proper-

    ties of PEI-impregnated on a variety of ordered mesoporous silicas.

    It was found that at constant PEI loading, the use of various sup-

    ports afforded different adsorption capacities, and that supported

    PEI had a higher capacity than its pure liquid counterpart. Interest-

    ingly, under otherwisethe sameconditions,the adsorption capacity

    appeared tobe dependenton pore diameter (dp).WhenPEI wasdis-

    persed on a KIT-6 type silica with 6nm dp  at a loading of 50%, thematerial adsorbed 3.07mmol/g in a stream of pure CO2   at 75

    ◦C,

    vs. 2.52mmol/g when using MCM-41 with 2.8nm pores as sup-

    port. Theadsorption capacityof 50%PEI-loaded KIT-6 in conditions

    closer to flue gas was 1.95mmol/g in the presence of 5% CO2/N2at 75 ◦C. The pore size can also affect the rate of adsorption as the

    time required toachieve90%of the total capacitywasin theorder of 

    KIT-6 < SBA-16= SBA-15< MCM-48< MCM-41. Following the ratio-

    nale that using large pore sizes afford better adsorption capacity,

    PEI and tetraethylenepentamine (TEPA) were impregnated on a sil-

    ica monolith with hierarchical pore structure [96]. Due to its largerdp, with mean values at 3, 17 and 120n m, the optimum loading

    of PEI was 65wt%, with a capacity of 3.75mmol/g for a stream of 

    5% CO2  in N2  at 75◦C, a capacity much higher than that obtained

    using conventional MCM-41 mesoporous silica as support. TEPA-

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    7/15

    766   A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760–774

    loaded samples were not particularly attractive as their adsorption

    capacity deteriorated after only 5 adsorption–desorption cycles.

    Anotherapproachusedby Yueet al.[97] consistedof impregnat-

    ingTEPA on as-synthesizedSBA-15,as opposed to calcinedsupport.

    The reported adsorption capacity for a sample with 50wt% load-

    ing in the presence of 10% CO2  in N2  was ca. 3.25mmol/g at 75◦C.

    The materials prepared with as-synthesized support consistently

    performed better thanthose prepared usingthe correspondingcal-

    cined supports, with up to 10% higher capacity with the added

    advantage that no steps are required to remove the organic tem-

    plate. The proposed explanation was that the polymeric template

    in the pores of as-synthesized supports interacts with the TEPA,

    forming a more even distributionof the functional groups, andpre-

    venting TEPA from aggregating into a micellar-like form, which

    is believed to be its naturally occurring form. The use of as-

    synthesized supports was further explored by Yue et al. [98] using

    MCM-41-type silica. A TEPA-MCM-41 sample loaded with 50 wt%

    PEI had a capacity of 4.54mmol/g for 5% CO2/N2  at 75◦C, outper-

    forming the above-mentioned TEPA-SBA-15.

    In addition, Yue et al. [99] impregnated as-synthesized SBA-15

    using a mixture of TEPA anddiethanolamine(DEA). In this case, the

    maximum CO2/N ratio (adsorption efficiency) was found to be ca.

    0.4 at a loading of ca. 30% TEPA and 20% DEA. In the case of TEPA-

    DEA-SBA-15 at 75◦C, its adsorption capacity ranged from 3.77 to3.61mmol/gthroughout 6 adsorption–desorptioncycles for 5% CO2in N2. Thegoodperformanceof this adsorbentwas attributed to the

    hydroxyl groups present in DEA. This is analogous to the effect of 

    water vapor, which is associated with a more favorable CO2  to N

    stoichiometry, as shown in Scheme 1.

    Franchi et al. [100] impregnated DEA on a variety of sup-

    ports whose adsorption capacity and stability were compared to

    a benchmark adsorbent, i.e., 13X zeolite. The most promising sup-

    port used in this work consisted of MCM-41 silica with pores

    enlarged by post-synthesis treatment (PE-MCM-41, dp =9.7nm),

    which afforded an adsorption capacity of 3 mmol/g at 25◦C in the

    presence of 10% CO2   in N2, a value comparatively higher than

    that reported for 13X (i.e., 2.8 mmolg−1) under the same condi-

    tions. Similarly to supported PEI, the adsorption capacity increasedwith DEAloading, butthe adsorption efficiency (CO2/N) decreased,

    suggesting that loading more than 6mmol of DEA per gram of  

    adsorbent is unattractive.

    A series of amine-impregnated mesoporous aluminas (MA),

    using diisopropanolamine, triethanolamine, 2-amino-2-methyl-

    1,3-propanediol, diethylenetriamine (DETA) and PEI, were inves-

    tigated by Plaza et al. [101]. The most attractive materials were

    those containing PEI and DETA with 40wt% loading. Not only did

    these adsorbents offer higher adsorption capacity at room temper-

    ature but, unlike the other materials in this study, their capacity

    increased at higher temperature. DETA-MA exhibited a capacity of 

    ca. 1mmol/g at 25 ◦C for pure CO2, increasing to ca. 1.4 mmol/g

    at 57 ◦C. For PEI-MA, the adsorption capacity at 57 ◦C was ca.

    1.14mmol/g in the presence of pure CO2. It is worth noting thatthe adsorption capacity was enhanced at higher temperature only

    for samples with high amine loadings, supporting the hypothesis

    that this behavior is associated with diffusion limitations within

    the amine phase at low temperature.

    A number of literaturereports explored the regeneration behav-

    ior of PEI-impregnated mesoporous materials. Drage et al. [11] used

    a proprietary mesoporous silica impregnated with PEI (40 wt%).

    The adsorbent showed a capacity of 2.4mmol/g at 70 ◦C in the

    presence of 15%CO2 in N2. This work analyzed the effectof regener-

    ation temperature using pure CO2 as stripping gas. It was observed

    that desorption at a temperature of less than 140 ◦C resulted in

    an incomplete regeneration. However, some concerns were raised

    regarding the use of such high temperatures, mainly because of 

    the following problems: (i) evaporation of PEI may occur and (ii)

    a secondary reaction between CO2   and amine groups formed a

    stable product, most likely urea, resulting in a decreasing num-

    ber of adsorption sites. The suggested alternative was to use a

    different stripping gas or lower desorption temperature although

    sacrificing some working adsorption capacity. As discussed later, a

    strategy to prevent the formation of urea during extensive cycling

    even at high temperature has been proposed recently by Sayari

    and Belmabkhout [102]. A PSA strategy was explored by Dasgupta

    etal. [103] using50% PEI-impregnatedSBA-15. The highestcapacity

    reported at 75◦C was 1.36 mmol/g for 12% CO2 in N2. A steadystate

    was obtained after 15–20 cycles, and the productivity was better

    compared to similar PSA procedure using 13X zeolite at 75◦C.

     3.1.2. Ordered microporous supports

    In addition to mesoporous materials, zeolites have also been

    used as supports. Jadhav et al. [104] dispersed monoethanolamine

    (MEA) on 13X zeolite producing materials with different loadings.

    Quite interestingly, the adsorbent with the highest capacity at low

    temperature (i.e., 35◦C), with 1.96mmolg−1 for 15% CO2   in N2,

    contained only 2.9wt% MEA, while the best capacity at 75◦C (i.e.,

    0.45mmol/g) was obtained on a sample with the highest load-

    ing (i.e., 25wt%). These capacities were comparatively higher than

    unmodified 13X, which adsorbed 0.64 and 0.36mmol/g at 35 and

    75 ◦C, respectively. An interesting advantage of amine-containing13X was a significant improvement in its tolerance to moisture.

    While it is generally accepted that preferential adsorption of water

    on 13X results in a drastic reduction of CO2  uptake, the adsorption

    capacity in the presence of 100% RH decreased by only ca.13% with

    respect to dry conditions.

    Another type of zeolite, namely beta-zeolite, wasused by Fisher

    et al. [105] to support TEPA and compared with TEPA-impregnated

    on amorphous alumina and silica. The results clearly showed the

    advantages of usinga support withgood structural propertiessince

    beta-zeolite was loaded with up to 38.4wt% compared to only 14.6

    and 8.3wt% for SiO2 and Al2O3, respectively, most likely as a result

    of a comparatively higherpore volume. Such loading translated in a

    significantly higher adsorption capacity for TEPA-beta zeolite over

    the other samples, being 2.08 mmol/g for 10% CO2  balance nitrogenat 30 ◦C, while it was 0.19 and 0.68mmol/g for TEPA-Al2O3   and

    TEPA-SiO2 , respectively.

     3.1.3. Other supports

    While ordered mesoporous supports are suitable substrates

    for the dispersion of amines, other supports were also explored.

    Extensive work performed by Filburn’s group [106–108], dealt

    with impregnation of a variety of amines, such as PEI,

    monoethanolamine, diethanolamine, triethanolamine, and TEPA

    on high surface area polymeric supports, mainly polymethyl-

    methacrylate (PMMA). Although their original purpose was to

    produce adsorbents for air purification in confined environments,

    the results have proven to be of interest for other applications

    such as flue gas treatment as shown in a later contribution [108],where a TEPA-impregnated PMMA exhibited capacities of 21.45

    and 13.88mmol/g at 20 and 70◦C, respectively in the presence of 

    15% CO2  and 2.6% H2O balance N2. It is also worth noting that the

    reported adsorption capacities were remarkably higher than any

    other data reported in the literature. Moreover, contrary to amine-

    impregnated mesoporous inorganic supports, adsorption capacity

    decreased at higher temperature for the polymer-based adsor-

    bents. In the same work, TEPA was reacted with acrylonitrile to

    selectively transformprimaryaminesinto secondaryaminesbefore

    impregnation, under the premise thatsecondary amines are advan-

    tageous. However, this wasnot thecase,since the sampleproduced

    byimpregnation of themodifiedTEPA, referred to as TEPAN,under-

    performed TEPA-PMMA in terms of adsorption capacity and rate

    with adsorption capacity valuesof 14.22 and 4.01 mmol/g at 20 and

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    8/15

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    9/15

    768   A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760–774

    Fig. 5. Amine loading (left) and adsorption capacity (right) vs. TRI/SiO2 ratio on MCM-41 (TRI-M41C)and PE-MCM-41 (TRI-M41EC)[117].

    favorable stoichiometry is expected in the presence of moisture,

    which was corroborated by the experimental value of adsorption

    capacity, increasing from0.41 to 0.89mmol/gin a wet stream (100%

    RH). For several years thereafter, no further CO2 adsorption studies

    on amine-grafted materials appeared, most likely because of the

    lack of interest in this topic. However, with the rapid development

    of ordered mesoporous materials, and the increasing awareness

    of the greenhouse gas effect, new studies began to appear in the

    literature starting in 2003.

    In a comparative investigation on propylamine-grafted MCM-

    48 and silica xerogel, Huang et al. [114] provided evidence for the

    advantages of ordered mesoporous supports. Using 10% CO2 in N2,

    a significantly higher capacity of ca. 1.42 mmol/g was obtained forAP-MCM-48 at room temperature vs. ca. 0.58mmol/g for amine-

    grafted silica xerogel under the same conditions. Since the amine

    loading was 2.3 and 1.7mmol/g for MCM-48 and silica xerogel,

    respectively, it is inferred that the significantly higher capacity

    of AP-MCM-48 was accompanied by an improved efficiency. Fur-

    ther, using a CO2-containing stream with 100% RH gave rise to an

    adsorption capacity twice as high with a CO2/N = 1 corresponding

    to quantitative transformation of amine groups into ammonium

    bicarbonate.

    Further evidence for the suitability of periodic mesoporous sup-

    ports was reported by Knowles et al. [115] using AP-grafted hexag-

    onal mesoporous silica (HMS) and amorphous silica gel. A higher

    amine loading of 2.3mmol/g was obtained vs. only 1.1mmol/g

    for amorphous silica. These loadings mirrored the difference insurface areas of HMS (1198 m2/g) compared to amorphous silica

    (567m2/g). The adsorption capacity at 20◦C inthe presenceof 90%

    CO2/Ar was 1.59mmol/g for AP-HMS compared to 0.68mmol/g

    for AP-grafted amorphous support. In further work, Knowles

    et al.[116] used (3-trimethoxysilylpropyl)diethylenetriamine (TRI)

    grafted on HMS and obtained a capacity of 1.34mmol/g in the

    presence of 90% CO2   balance Ar at 20◦C. They also found that the

    material is thermally stable up to 170 ◦C under pure N2   or mildly

    oxygenated environments, a comparatively higher temperature

    than amine-impregnated adsorbents.

    Sayari’s group made significant contributions to the area of CO2capture by amine-containing nanoporous materials. They demon-

    strated the beneficial effect of using materials with larger pore

    diameterand porevolume thantypical MCM-41 silica [90,117,118].

    To do so, they used the post-synthesis pore-expansion method

    developed earlier [119,120]. Based on as-synthesized MCM-41 as

    starting material, they generated PE-MCM-41 with pore size and

    pore volume up to 20n m and 3.5c m3/g, vs. typically ca. 3–4nm

    and ca. 0.7–1cm3/g for regular MCM-41, with hardly any change

    in surface area. As shown in Fig. 5, grafting MCM-41 and PE-MCM-

    41 with TRI led to comparable amine loadings, because of similar

    surface areas. However, as shown in Fig. 5, using 5% CO2   in N2at 25 ◦C, the CO2   uptake was ca. 50% higher for TRI-PE-MCM-41

    than TRI-MCM-41, at all amine loadings. Moreover, TRI-PE-MCM-

    41 adsorbed CO2   about 30% faster than MCM-41-based material,

    showing the importance of pore size and volume.

    Another contribution of Sayari’s group was the optimizationof the grafting conditions, leading to dramatic improvement of 

    amine loading and adsorptive properties. Grafting is traditionally

    practiced under reflux, in dry solvent (typically toluene at 110 ◦C)

    with large excess of silane. Harlick and Sayari [90] found that

    the optimum grafting conditions of TRI on PE-MCM-41 in toluene

    were as follows: T = 85 ◦C; water added: 0.3mL per gram of sup-

    port; aminosilane added: 3 mL per gram of support. Under such

    conditions, the amine content increased by ca. 30% (i.e., 7.98 vs.

    6.11mmol/g for conventional dry grafting), whereas the adsorp-

    tion capacity using 5% CO2/N2   at 25◦C increased by ca. 70% from

    1.55mmol/g for conventional dry grafting to 2.65mmol/g. Thus,

    under these CO2   adsorption conditions, the combination of pore

    expansion and optimization of grafting conditions improved the

    adsorption capacity by close to 300% compared to the adsorbentproducedvia anhydrousgrafting on conventional MCM-41, in addi-

    tionto a significant increasein therate of adsorption. The advantage

    of using amine-functionalized mesoporous materials was further

    evidenced when a stream of humid CO2  w as used. In the pres-

    ence of 5% CO2   in N2   with 27% RH, the adsorption capacity for

    TRI-PE-MCM-41 increasedto 2.94 mmol/gin contrast to a dramatic

    decrease observed for 13X, down to 0.09 mmol g−1. This work also

    provided evidence of the advantage of amine-grafted adsorbents

    in terms of adsorption kinetics, as the CO2   rate of adsorption on

    TRI-PE-MCM-41 was found to be higher than 13X zeolite. Later, it

    would be corroborated that TRI-PE-MCM-41 is also comparatively

    faster than its PEI-impregnated counterpart [91].

    Further studies on TRI-PE-MCM-41 [118] demonstrated that

    enhanced capacity was not the only advantage of TRI-PE-MCM-

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    10/15

     A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760–774 769

    Fig. 6. Working adsorption capacity of TRI-PE-MCM-41 over various adsorption–desorption cycles in dry (TRI-70/70-d) and humid (TRI-70/70-h) streams with adsorption

    and desorption at 70 ◦C [102].

    41, since incorporation of amines significantly increased the

    selectivity toward CO2   over N2. Using conditions directly related

    to flue gas, i.e., 10% CO2   balance N2   at 50◦C, Serna-Guerrero

    et al. [121] obtained a stable capacity of 1.59mmol/g over 100adsorption–desorption cycles with regeneration under vacuum at

    90 ◦C.

    To further address the long-term stability of amine-

    functionalized adsorbents for CO2 capture,Sayari andBelmabkhout

    [102] carried an in-depth investigation using extensive

    adsorption–desorption cycling under different conditions. As

    shown in Fig. 6, they found that under dry conditions, the adsor-

    bent will ultimately deactivate even under mild conditions. The

    degree of deactivation depended on the nature of the adsorbent

    and the adsorption–desorption conditions. The adsorbent deacti-

    vation was clearly associated with the formation of urea groups,

    which are stable under the desorption conditions. To prevent the

    formation of urea and drastically improve the stability of amine-

    functionalized adsorbents, the use of humid streams was proposed.As illustrated in Fig. 6, the adsorbent underwent more than 700

    cycles without any loss in adsorption capacitywhen the adsorption

    anddesorption gases contained7% RH at 70◦C. Another interesting

    finding was that by treating deactivated AP-grafted PE-MCM-41

    in the presence of water vapor at ca. 200 ◦C, it was possible to

    hydrolyze the urea groups and fully regenerate the adsorbent.

    The advantages of using TRI were also discussed by Hiyoshi

    et al. [122] in a thorough comparative analysis of monoamine,

    diamine and triamine-bearing molecules grafted on SBA-15. The

    higher amine density achieved through the use of TRI resulted in

    the best performing adsorbent. The reported adsorption capacity

    for TRI-SBA-15 at 60◦C and 15% CO2 was of 1.58mmol/g under dry

    conditions and 1.80mmol/g in a stream containing 60% RH. Fur-

    thermore, they reported that TRI-SBA-15 was stable over 50 cyclesof adsorption at 60◦C and desorption at 100 ◦C.

    Kimetal. [123] made a comparative studybetweenmesoporous

    silica (MS) functionalized with molecules containing 1–3 amine

    groups produced by anhydrous grafting and co-condensation.

    In general, samples prepared by co-condensation presented

    higher amine contents, that reportedly promoted a better amine-

    efficiency and adsorption capacity. In line with the observations

    mentioned above, the adsorbent with the highest capacity was

    TRI-MS, with 1.74mmol/g for pure CO2   at 25◦C, while the most

    efficient under the same conditions was the AP-MS sample with

    a CO2/N ratio of 0.43 and a capacity of 1.14mmol/g. This is con-

    sistent with findings by Serna-Guerrero et al. [124] who obtained

    the maximum efficiency of CO2/N= 0.5 using AP-grafted PE-MCM-

    41, whereas the CO2

    /N ratio for TRI-PE-MCM-41 never exceeded

    0.34 [121]. In addition, Kim et al. [123] compared their amine-

    grafted materials with PEI-impregnated KIT-6 silica. Although a

    higher capacity was obtained on the PEI-containing sample (i.e.,

    1.79mmol/g), its CO2/N efficiency at room temperature was only0.1.

    The use of diamine-bearing molecules was investigated under

    the hypothesis that the occurrence of two amine groups in

    close proximity will lead to enhanced formation of carbamate,

    thus higher CO2/N efficiency. Knofel et al. [125] grafted N -[3-

    (trimethoxysilyl)propyl] ethylenediamine (EDA) on SBA-16 silica.

    Although this work was mainly focused on CO2 adsorption at high

    pressure, it clearly showed that incorporation of amine groups

    resulted in an improved capacity at CO2   partial pressures below

    1 bar. Thereportedcapacityfor pure CO2 at 1barwas of 1.4mmol/g

    at27 ◦C for the best performing EDA-SBA-16. It was observed how-

    ever, that at high pressure (ca. 4 bar or more), the non-aminated

    samples exhibited higher adsorption capacity. A possible explana-

    tion was that physical adsorption predominates at high pressureand so, the higher pore volume of the unmodified support offers a

    comparative advantage in terms of adsorption capacity.

    In recent years, efforts to further improve the grafting pro-

    cess have been pursued, with the aim of improving the efficiency

    and capacity of aminated silicas. Wang et al. [126] incorporated

    AP-functionality by simultaneous extraction of structure direct-

    ing agent and grafting on as-synthesized SBA-15. The adsorbent

    obtained by the proposed approach outperformed a sample syn-

    thesized using the typical grafting procedure on calcined SBA-15.

    The sample usingas-synthesized support produced a material with

    an adsorption capacity of ca. 0.45mmol/g at 65◦C a t a C O2  partial

    pressure of 0.1bar, representinga CO2/N efficiency of 0.44, close to

    thestoichiometricratio of 0.5. It was suggestedthat, unlike calcina-

    tion, the extraction of surfactant template performed with ethanolpreserved the surface silanol groups, whichtranslated into a better

    distribution of surface amines with a subsequent improvement of 

    adsorption capacity.

    The drawback of surface silanol groups removal during calci-

    nation of the support was also addressed by Wei et al. [127]. They

    proposed rehydrating SBA-15by soaking it in water at 97◦C, before

    grafting with EDA. The obtained material had an amine content

    of 3.06mmol/g, and a capacity of 0.73mmolg−1 for 0.15bar CO2at 60 ◦C. A similar material prepared using non-hydrated SBA-15

    hadan amine loading of 2.59 mmol/g and an adsorption capacity of 

    0.59mmol/g.

    Zelenak et al. reported on the effect of pore size [128] and

    the basicity of the functional groups [129] on the performance of 

    amine-functionalized adsorbents for CO2

     capture. It was suggested

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    11/15

    770   A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760–774

    Fig. 7. Schematic representation of thesynthesis of hyperbranched aminosilica according to Jones et al. [136].

    that large pore sizes are associated with an efficient use of amine

    groups. Indeed, when grafting AP on MCM-41 with a pore size of 

    3.3nm, a high amine content was obtained (i.e., 3 mmol/g), but theadsorption capacity was only 0.57mmol/g for 10% CO2   at 25

    ◦C.

    In contrast, a capacity of 1.54mmol/g was obtained when SBA-15

    with a pore size of 7.1nm was used as support, despite a slightly

    lower amine loading of 2.7 mmol/g. The lower efficiency of MCM-

    41-basedmaterial, however,may notbe only a resultof a difference

    in pore sizes. The amine surface density reported for AP-MCM-

    41 was lower, having only 1.1 amine group per nm2 compared to

    2.4 amines per nm2 for AP-SBA-15. Since admittedly two amine

    molecules in close proximity are required for reaction with CO2,

    this would be a disadvantage for the lower amine density sample.

    Some interesting studies were devoted to the effect of the sup-

    ports on the adsorbent performance. Knofel et al. [130] compared

    AP-grafted mesoporous silica (MS) and mesoporous titania (MT).

    The highest adsorption capacity of ca. 0.24mmol/g for 10% CO2at 30 ◦C was obtained with AP-MS. This is a low capacity com-

    pared to other materials reported in the literature, but the main

    finding of this work was that the properties of the support may

    influence the behavior of the functionalized adsorbent. While no

    interactions were detected between CO2  and the silica support,

    interactions occurred in the presence of MT. This is reflected in

    a higher capacity when expressed in terms of surface area, i.e.,

    1mol/m2 and 0.6mol/m2 for AP-MT and AP-MS, respectively.

    Another approachexploredby Lu et al. [131] wasthe useof particles

    with defined geometry, by grafting EDA on mesoporous spherical

    particles. The adsorption capacity at 60◦C in the presence of 10%

    CO2 in air was of 0.73 mmol/g. These adsorbents showed a remark-

    able stability when regenerated using a TSA procedure at 120 ◦C

    or under VSA, although their adsorption capacity decreased in thefirst VSA cycle. In addition, it was suggested that the combina-

    tion of heat and vacuum resulted in an improvement in desorption

    rate.

    Looking for an inexpensive source of silica, Bhagiyalakshmi

    et al. [132] grafted tris(2-aminoethyl)amine (TREN) and TEPA onto

    chloropropyl-modified mesoporous supports produced from rice

    husk. The highest capacities were obtained with TREN-grafted

    MCM-48 with values of ca. 1.59 and 1.36mmol/g at 25 and 50◦C,

    respectively in the presence of pure CO2.

    The only contribution dealing with amine-grafted zeolites used

    ITQ-6[133], which offers attractive characteristics suchas highcon-

    centration of surface silanol groups anda poresize in thenanometer

    range. The mostpromising adsorbenthad a capacityof 0.67mmol/g

    for 12% CO2 at 20

    C.

     3.3. Hyperbranched aminosilicas

    A different method of functionalization, introduced recently,consisted in iterative building of amine-containing dendrimers

    inside the porous supports. Liang et al. [134,135] produced highly

    branched dendrimers by step-wise reaction between diisopropy-

    lethylamineand cyanuric chloride inside the pores of SBA-15 [134]

    or mesocellular siliceous foams [135]. The optimum adsorbent pro-

    duced with this approach was obtained after 3 reaction steps, with

    a capacity of ca. 1 mmol/g for 90% CO2   in Ar at 20◦C. However, as

    a larger number of reaction steps were performed to obtain higher

    generation dendrimers, the adsorbent lost its structural properties,

    allegedly due to space limitations, which negatively impacted the

    adsorptive properties.

     Jones’ group [136,137] proposed an innovative amine polymer-

    ization approach inside SBA-15 channels with promising results.

    In this case, aziridine was polymerized by ring opening inside thepores of SBA-15 producing a “covalently tethered hyperbranched

    aminosilica”, as represented in Fig. 7. This material exhibited a

    capacityof 3.11 mmol/g under a flowof water saturated10%CO2/Ar

    at 25 ◦C. The CO2/N efficiency was as high as 0.44 at room tem-

    perature, close to the theoretical value of 0.5. With respect to its

    performance at 75◦C, and 10% CO2/Ar, the hyperbranched-SBA-

    15 was stable, presenting an average adsorption capacity of ca.

    1.98mmol/g over 12 cycles with regeneration at 130 ◦C. In a later

    contribution [138], it was shown that higher loading of hyper-

    branched amines afforded a better capacity. The best reported

    adsorbent had an amine loading of 9.78mmol/g and adsorbed ca.

    4mmol/g at 10% CO2/N2 at75◦C in the presence of humidity.

    As summarized in Table 5, Similarly to amine-impregnated

    adsorbents, the covalently bonded aminated adsorbents spanmaterials with a wide variety of characteristics and performances.

    However, a number of common advantages and limitations of 

    amine-graftedmaterials canbe outlined. Onlysupportsthat exhibit

    surface hydroxyl groups can be used to produce amine-grafted

    materials. It was observed that high amine loading is a result

    of high surface area and availability of surface silanol groups,

    but the efficient use of functional groups is observed mainly

    in supports with large pores. While the equilibrium adsorption

    capacities are certainly not as high as those reported with some

    amine-impregnated adsorbents, properly designed amine-grafted

    materials do not exhibit the strong diffusion limitations observed

    in impregnated adsorbents. Therefore highadsorption ratesare not

    restricted upon operating at high temperature. A particular advan-

    tageoffered by amine-graftedadsorbents is theirhigh stabilityover

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    12/15

     A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760–774 771

     Table 5

    Literature data on CO2 adsorption capacity of amine-grafted adsorbents.

    Support   Amine Capacity

    (mmol/g)

    Amine

    loading

    (mmol/g)

    CO2/N   Experimental conditions Reference

    CO2  concentration T (◦C)

    Silica gel AP 0.89 1.26 0.71 100% (100% RH) 50 [113]

    MCM-48 AP 2.3 2.3 1 10% (100% RH) 25 [114]

    HMS AP 1.59 2.29 0.69 90% 20 [115]

    HMS TRI 1.34 4.57 0.29 90% 20 [116]

    PE-MCM-41 TRI 1.59 7.9 0.20 10% 50 [121]

    SBA-15 TRI 1.80 5.80 0.31 15% (humid) 60 [122]

    MS TRI (co-cond) 1.74 5.18 0.34 100% 25 [123]

    SBA-16 EDA 1.4 0.76 1.84 100% 27 [125]

    SBA-15 AP 0.45 2.56 0.18 10% 65 [126]

    SBA-16 EDA 0.727 3.06 0.24 15% 60   [127]

    SBA-15 AP 1.54 2.72 0.57 10% 25 [128]

    SBA-12 AP 1.04 2.13 0.49 10% 25 [129]

    MS AP 0.24 1.6 0.15 10% 30 [130]

    MSP EDA 0.73 0.99 0.73 10% 60 [131]

    MCM-48 TREN 1.36 4 0.34 100% 50 [132]

    ITQ-6 AP 0.67 1.26 0.53 12% 20 [133]

    SBA-15 Amine-dendrimers 1 1.25 0.40 90% 20 [134]

    SBA-15 Aziridine polymer 4 9.78 0.41 10% (humid) 75 [138]

    hundreds, most likely thousands, of adsorption–desorption cycles

    [102].

    Although this review focused on CO2  capture, as mentioned in

    the Introduction, there are other impuritiesin flue gas. Of particular

    concern with respect to amine-functional materials is the pres-

    ence of SO2, as it was found that it affects negatively their cyclic

    performance. For example, it was recently reported thatafter expo-

    sure to SO2, the working adsorption capacity of TRI-PE-MCM-41

    decreased from 1.57mmol/g to 0.89mmol/g for a mixture of 10%

    CO2/N2 at50◦C [139]. Supportedby gravimetric measurementsand

    FTIRspectroscopy,it wasproposed thatSO2 reacts irreversibly with

    the primary amines of the triamine functional molecules. Conse-

    quently, it might be necessary to engineer processes to remove

    SO2   prior to CO2   capture to preventing its contact with amine-

    functionalized adsorbents.

    4. Conclusions

    Major advances have been achieved towardthe development of 

    a CO2 capture technology basedon adsorption. Physical adsorbents

    suchas zeolites,carbon-basedmaterials andMOFs werefound to be

    suitable,mostlyat lowtemperatureand highpressure. Theseadsor-

    bents, however, often adsorb water vapor preferentially over CO2 ,

    andtheir CO2 adsorption capacityat low pressure is not sufficiently

    high. Although these materials may provide elegant solutions for

    CO2 sequestration and storage, theyare not particularly suitable for

    post-combustion gas treatment. Nevertheless, a continuous effort

    is being deployed to circumvent such drawbacks. The strategies

    beingused include surface modificationto enhancethe interactions

    with CO2, thus increasing the adsorption capacity at low pressure.Another route is to design completely newmaterials such as ZMOFs

    andCOFs with increased tolerance to moisture in the gas feed, thus

    improved CO2 selectivity.

    Likewise, tremendous progress has been achieved in the devel-

    opment of novel chemical adsorbents such as amine-modified

    materials with large surface area. By optimizing the synthesis con-

    ditions and using supports with adequate structural properties, it

    was possible to develop materials with superior CO2  adsorptive

    properties, particularly suitable for flue gas treatment. Typically,

    these materials exhibit large CO2  adsorption capacity even at low

    pressure, high rate of adsorption anddesorption, andexcellent tol-

    erance to moisture in the feed. Furthermore, contrary to physical

    adsorbents, the selectivity of amine-functionalized materials is not

    significantly affected by temperature, at least within the range of 

    interest for flue gas treatment. While the stability of this kind of 

    adsorbents has been questioned, it was recently demonstrated that

    their stability may be dramatically enhanced during thousands of 

    adsorption–desorption cycles, provided that the feed and purge

    gases contains moisture. The role of moisture is to prevent the

    formation of urea linkages, which is the main source of material

    deactivation.

    This review clearly showed a steady improvement in the CO2adsorptive properties of novel materials. The course followed so

    far has resulted in major achievements that may well pave the way

    for an alternative CO2 capture technology in the near future.

    References

    [1] D.J. Hofmann,J.H.Butler,P.P.Tans,A newlookat atmosphericcarbondioxide,Atmos. Environ. 43 (2009) 2084–2086.

    [2] R. Monastersky,A burdenbeyond bearing, Nature 458 (2009) 1091–1094.[3] A.J. Yamakasi, An overview of CO2   mitigation options for global warming

    emphasizing CO2  sequestrationoptions,J. Chem.Eng.Jpn. 36 (2003)361–375.[4] D.Aaron,C. Tsouris,Separationof CO2 from fluegas: a review, Sep. Sci. Tech-

    nol. 40 (2005) 321–348.[5] S.M. Benson, F.M. Orr Jr., Carbon dioxide capture and storage, MRS Bull. 33

    (2008) 303–305.[6] K.M.K. Yu, I. Curcic, J. Gabriel, S.C.E. Tsang, Recent advances in CO2  capture

    and utilization, ChemSusChem 1 (2008) 893–899.[7] H. Audus, Greenehouse gas mitigation technology: an overview of the CO2

    capture andsequestrationstudiesand further activitiesof theIEA greenhousegas R&D programme, Energy22 (1997) 217–221.

    [8] A. Sjostrom, H. Krutka, Evaluation of solid sorbents as a retrofit technologyfor CO2  capture, Fuel 89 (2010) 1298–1306.

    [9] M.T. Ho, G.W. Allinson, D.E. Wiley, Reducing the cost of CO2  capture fromflue gases using pressure swing adsorption, Ind. Eng. Chem. Res. 47 (2008)4883–4890.

    [10] S. Majumdar, A. Sengupta, J.S. Cha, K.K. Sirkarst, Simultaneous SO2/NOx sep-aration from flue gas in a contained liquid membrane permeator, Ind. Eng.

    Chem. Res. 33 (1994) 667–675.[ 11] T.C. Dra ge, A. A renillas, K.M. Smith, C .E. Snape, Thermal stability of  

    polyethylenimine based carbon dioxide adsorbents and its influence onselection of regeneration strategies, Micropor. Mesopor. Mater. 116 (2008)504–512.

    [12] S.Choi,J.H.Drese, C.W. Jones,Adsorbentmaterials forcarbondioxide capturefrom large anthropogenic point sources, ChemSusChem 2 (2009) 796–854.

    [13] K.B. Lee, M.G. Beaver, H.S. Caram, S. Sircar, Reversible chemisorbents for car-bon dioxide and their potential applications, Ind. Eng. Chem. Res. 47 (2008)8048–8062.

    [14] C.J. Major, B.J. Sollami, K. Kammermeyer,Carbon dioxide removal from airbyadsorbents, Ind. Eng. Chem. Process Des. Dev. 4 (1965) 327–333.

    [15] D.M. Ruthven, Principles of Adsorption and Adsorption Processes, Wiley-Interscience, New York, 1984.

    [16] R.T. Yang, Gas Separation by Adsorption Processes, Imperial College Press,Boston, 1997.

    [17] D.Lozano-Castello, D. Cazorla-Amoros,A. Linares-Solano, Powderedactivatedcarbonsand activatedcarbonfibersfor methanestorage:a comparativestudy,

    Energy Fuels 16 (2002) 1321–1328.

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    13/15

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    14/15

  • 8/9/2019 Flue gas treatment via CO2 adsorption.pdf

    15/15

    774   A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760–774

    [135] Z. Liang,B. Fadhel, C.J. Schneider, A.L. Chaffee, Adsorption of CO2 on mesocel-lular siliceous foamiteratively functionalized with dendimers, Adsorption15(2009) 429–437.

    [136] C.W. Jones, J.C. Hicks,D.J. Fauth,M. Gray,Structures for capturing CO2, meth-o ds o f making t he st ructure s and m et hods o f capturing CO2 , US PatentApplication No. US2007/0149398, 2007.

    [137] J.C. Hicks, J.D. Drese, D.J. Fauth, M.L. Gray, G. Qi, C.W. Jones, Designing adsor-bents forCO2 capture from flue gas – hyperbranched aminosilicas capable of capturing CO2 reversibly, J. Am. Chem. Soc. 130 (2008) 2902–2903.

    [138] J.H. Drese, S. Choi, R.P. Lively, W.J. Koros, D.J. Fauth, M.L. Gray, C.W. Jones,Synthesis–structure–property relationships for hyperbranched aminosilicaCO2 adsorbents, Adv. Func. Mater.19 (2009) 3821–3832.

    [139] Y. Belmabkhout, A. Sayari, Isothermal versus non-isothermaladsorption–desorption cycling of triamine-grafted pore-expanded MCM-41mesoporous silica for CO2 capture from flue gas, Energy Fuels 24 (2010)5273–5280.