ultrafiltration membrane synthesis by nanoscale templating

Upload: riki-mandol

Post on 14-Apr-2018

243 views

Category:

Documents


1 download

TRANSCRIPT

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    1/14

    Journal of Membrane Science 198 (2002) 173186

    Ultrafiltration membrane synthesis by nanoscale templating ofporous carbon

    Michael S. Strano a,1, Andrew L. Zydney a,1, Howard Barth b,Gilber Wooler b, Hans Agarwal a,1, Henry C. Foley c,

    a Colburn Laboratory, Department of Chemical Engineering, University of Delaware, Academy Street, Newark, DE 19716, USAb DuPont Company, Central Research and Development, Experimental Station, P.O. Box 80228, Wilmington, DE 19880-0228, USA

    c Department of Chemical Engineering, Center for Catalytic Science and Technology,

    Pennsylvania State University, University Park, PA 16802, USA

    Received 22 March 2001; received in revised form 28 June 2001; accepted 9 July 2001

    Abstract

    A novel method for producing carbon membranes for ultrafiltration applications is presented using a spray deposition and

    pyrolysis of poly(furfuryl alcohol)/poly(ethylene glycol) mixtures on macroporous stainless steel supports. The poly(ethylene

    glycol) or PEG employed as a carbonization template creates a mesoporosity that leads to pores in the ultrafiltration range.

    Scanning electron microscopy (SEM) shows that the membranes consisted of 12- to 15-m thick carbon films. Gas permeation

    and water permeability data were used for the calculation of mean pore sizes, which were found to decrease with decreasingaverage molecular weight of the PEG template. Ultrafiltration of a polydisperse dextran solution was used to quantify the

    retention properties of the membranes. Molecular weight cutoffs determined from dextran retention data were shown to vary

    with template molecular weight: values of 2 104, 3.5 104, and 6 104 gmol1 dextran were measured for respective

    templates of 2000, 3400, and 8000g mol1 PEG. For PEG molecular weights of 2000 or below, the templating effect was

    ill defined, membrane film cracking became more prominent, and membrane selectivity and reproducibility were adversely

    affected. 2002 Published by Elsevier Science B.V.

    Keywords: Ultrafiltration; Carbon membranes; Membrane formation; Inorganic membranes

    1. Introduction

    Membrane filtration technologies are critical to a

    variety of industrial process applications including

    cell harvesting, sterile filtration, protein enrichment,

    and removal of particulate matter [1]. Ultrafiltration

    Corresponding author. Tel.: +1-814-865-2574;

    fax: +1-814-865-7846.

    E-mail addresses: [email protected] (M.S. Strano),

    [email protected] (H.C. Foley).1 Tel.: +1-302-831-2345; fax: +1-302-831-2085.

    in particular, where the membrane retains compo-

    nents with kinetic diameters from 1 to 100 nm, haswidespread utility for industrial separation processes.

    Traditionally, membranes used for ultrafiltration have

    been polymeric in nature. Asymmetric ultrafiltration

    membranes are commonly synthesized using phase

    inversion, where a polymer solution consisting of

    a base and pore-former in a solvent is induced to

    form two interdispersed liquid phases. Membranes

    synthesized in this manner include the bilayer type

    containing slit shaped fissures or cracks [2] and those

    that contain plasticizers and are stable while dry [3,4].

    0376-7388/02/$ see front matter 2002 Published by Elsevier Science B.V.

    P I I : S 0 3 7 6 - 7 3 8 8 ( 0 1 ) 0 0 5 7 4 - 9

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    2/14

    174 M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186

    Nomenclature

    B0 porous structure factor

    (mol ideal gas m

    1

    )Cb concentration of dextran in

    bulk solution (g l1)Cp concentration of dextran in

    permeate (g l1)

    Cw concentration of dextran at

    membrane surface (g l1)dm average membrane pore size from

    gas permeation (m)

    J flux of gas (mol m2 s1)

    Jv filtrate flux during ultrafiltration (m s1)

    k solute mass transfer coefficient (m s1)K0 Knudsen permeability

    (mol m1 s1 Pa1)

    l carbon layer thickness (m)

    M molecular weight (g mol1)p average trans-membrane pressure

    difference (Pa)

    p pressure driving force under gas

    permeation (Pa)

    rpore cylindrical pore radius (m)

    R ideal gas constant (Pa m3 mol1)

    Rc rejection coefficient for dextran

    ultrafiltrationT temperature of gas permeation (K)

    Greek symbols

    geometric constant for porous materials

    geometric constant for porous materials

    proportionality constant in mass

    transfer correlation

    stagnant film thickness above

    membrane in stirred-cell (m) membrane porosity

    gas viscosity (Pa s) viscosity of solvent (Pa s)

    pore tortuosity

    Despite the widespread use of these types of

    membrane materials, they suffer from several disad-

    vantages. The membranes have limited mechanical

    integrity which can lead to deformation during op-

    eration and adversely affect membrane performance.

    These materials often have limited temperature and

    chemical stability, prohibiting their use in applica-

    tions involving harsh organic solvents or at elevated

    temperature [4]. The use of caustic and hypochlorite

    solutions as cleaning agents can also lead to chemicaldegradation of the polymeric materials, significantly

    reducing the life of these membranes. In addition,

    membranes made from hydrophobic polymers like

    polysulfone and polyethersulfone must contain a

    humectant such as glycerol to avoid problems asso-

    ciated with re-wetting, while membranes made from

    hydrophilic polymers like cellulose must remain wet

    at all times to avoid collapse of the pore structure.

    While inorganic ultrafiltration membranes have

    attracted increasing attention in recent years [57],

    pore size modified carbon membranes have been

    relatively unexplored. Yet compared to ceramic and

    metallic ultrafiltration membrane materials, carbon

    can be considerably less expensive. Nanoporous car-

    bon is also a promising material because it is chem-

    ically inert under typical processing conditions and

    thermally stable at temperatures well above 200 C

    [8,9]. Formed from the pyrolysis of carbonizing nat-

    ural or synthetic polymeric precursors, this material

    is a disordered solid having a pore size approach-

    ing molecular dimensions and has been shown to

    possess highly shape-selective molecular transport

    properties [10]. Poly(furfuryl alcohol) (PFA)-derivednanoporous carbons have a mean pore size about

    0.5 nm as measured from N2 and methyl chloride ad-

    sorption isotherms [11]. Attempts to use this material

    in the synthesis of defect free, micron scale films on

    structurally stable macroporous supports have been

    successful. Nanoporous carbon membranes have been

    fabricated with high selectivity for some gas phase

    separations of small molecules [1214].

    While such membranes have a mean pore size on

    the molecular scale (

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    3/14

    M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186 175

    and 0.28 cm3 g1 for 2000, 3400, and 8000 g mol1

    templates, respectively. These results have been used

    in the synthesis of various nanoporous carbon cat-

    alysts with greatly enhanced effectiveness factorswhen compared to catalysts prepared from native

    nanoporous carbon alone [16,17].

    Studies involving the use of porous carbon for ul-

    trafiltration membrane formation in the literature are

    scarce. Schindler and Maier [4] claimed to have syn-

    thesized such a membrane through the carbonization

    of a pre-existing polymer membrane with identical

    pore structure. Although pyrolysis of a polymeric

    microfiltration membrane can reasonably produce a

    carbon membrane with similar macroporosity, there is

    little theoretical or empirical evidence to suggest that

    pores in the entire ultrafiltration range (1100 nm)

    would be preserved during the carbonization of a

    pre-existing film. Additionally, ultrafiltration range

    pores were undetectable using the bubble-point test-

    ing methodology employed by the authors [4].

    In this paper, a novel method for producing carbon

    ultrafiltration membranes is presented. Spray deposi-

    tion and pyrolysis on macroporous stainless steel of

    poly(furfuryl alcohol)/poly(ethylene glycol) mixtures

    is the straightforward fabrication technique employed.

    The methodology is demonstrated in Fig. 1. As in our

    earlier work, the PEG is used as a non-carbonizingtemplate molecule to introduce controllable meso-

    porosity into the carbon film. Permeation of a poly-

    disperse dextran solution was used to quantify the

    Fig. 1. Mesoporous carbon membranes via nanoscale templating.

    retention properties of the membranes as a function

    of template size. A correlation of partial rejection

    coefficients for 90% of the dextran mixture with tem-

    plate molecular weight was observed. Membraneswere also characterized using scanning electron

    microscopy (SEM) and gas permeation.

    2. Experimental

    2.1. Membrane synthesis

    Carbon based ultrafiltration membranes were syn-

    thesized using an adaptation of a spray coating proce-

    dure used for creating supported nanoporous carbon

    membranes [13]. Furfuryl alcohol resin in the form of

    Durez Resin #16,470 (PFA) from Occidental Chem-

    ical Corporation diluted in reagent grade acetone

    (Aldrich Chemical) was combined with poly(ethylene

    glycol) (PEG) and spray deposited upon a macrop-

    orous stainless steel support. The PFA resin has a

    specific gravity of 1.21 and a viscosity of 200 cP

    at 25 C that decreases to about 5 cP at 80 C. PEG

    samples of average molecular weights of 1000, 1500,

    2000, 3400, and 8000 g mol1 were obtained from

    Aldrich Chemical Company.

    This PFA/PEG/acetone precursor was prepared bycombining PFA resin with a given average molecular

    weight of PEG in a 50/50 wt.% ratio at 70 C and sub-

    sequent dilution in acetone forming a stable mixture

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    4/14

    176 M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186

    Table 1

    Summary of synthesis and performance characteristics for membranes used in this worka

    Initial weight

    support (g)

    Wet weight

    support (g)

    Net carbon

    weight (g)

    Pressure

    applied (psig)

    Water flux

    (m s

    1

    )

    Water permeance

    (m s

    1

    Pa

    1

    )M8000-1-0.021 10.5107 10.6393 0.021 50.5 1.8 106 5.1 1012

    M8000-2-0.018 10.8591 10.9768 0.018 54 1.1 105 2.9 1011

    M8000-3-0.022 10.4478 10.58 0.022 54 1.8 106 4.7 1012

    M1500-1-0.019 10.4659 10.5728 0.019 55 5.9 107 1.5 1012

    M3400-1-0.019 10.7327 10.8409 0.0188 75 8.3 107 1.6 1012

    M3400-2-0.016 10.286 10.3696 0.0161 75 8.3 107 1.6 1012

    M3400-3-0.018 10.5444 10.6481 0.018 75 8.3 107 1.6 1012

    M2000-1-0.019 10.6382 10.7455 0.0189 50 2.0 108 5.6 1014

    M2000-2-0.019 10.4353 10.534 0.019 50 1.3 107 3.8 1013

    M2000-3-0.019 10.3555 10.4623 0.019 50 8.8 108 2.5 1013

    M2000-4-0.019 10.6816 10.7785 0.0189 50 1.0 107 3.0 1013

    M1000-1-0.029 10.5297 10.6695 0.0288 50 1.7 107 5.0 1013

    M1000-2-0.023 10.7057 10.8189 0.0227 50 5.7 108 1.6 1013

    a Labels are as follows: M(PEG MW)-(sample #)-(carbon mass (in g)).

    at room temperature with a viscosity near 5 cP. The

    resulting solution was subsequently spray deposited

    onto circular porous stainless steel supports (0.2m

    pore size, 11.4 cm2 area, and 1 mm thickness) sup-

    plied by Mott Metallurgical Co. Before spray coating

    the stainless steel supports were cleaned by sonica-

    tion in acetone. After spray coating approximately

    200 mg of the polymer blend solution (wet weight)was deposited on the support. The resulting system

    was subsequently pyrolyzed in a stream of flowing

    He (50sccm) at 5 Cmin1 to 600 C, held for 2h

    at this temperature, and then allowed to cool to room

    temperature. The final carbon mass on each support

    was approximately 20 mg. Table 1 presents data on

    the membranes used in this work. Membrane labels

    contain the PEG template molecular weight and total

    coat mass of carbon for each sample.

    2.2. Scanning electron microscopy

    SEM was used to characterize the internal mor-

    phology of the membranes. Membrane cross-sections

    were cut orthogonal to the support surface using a

    diamond-wafering saw. These sections were mounted

    in an epoxy resin, polished, and given a coating of

    Au for imaging with a Hitachi S-4000 field emis-

    sion scanning electron microscope. Imaging was

    performed on areas of these cross-sections external

    to and within the macroporosity of the stainless steel

    support and at various radii from the center of the

    disk-shaped membranes.

    2.3. Characterization using gas permeation

    Gas phase transport of He, Ar, N2, O2, SF6 and CO2was used to characterize the selective porosity and

    integrity of the carbon membrane. The disk-shapedmembranes were sealed using VitonTM gaskets into a

    stainless steel module setup to measure the transport

    of a single gas through the membrane. The pure com-

    ponent flux of each gas was measured as a function

    of pressure. Details concerning the experimental setup

    and technique can be found in [18].

    2.4. Ultrafiltration of polydisperse dextrans

    Ultrafiltration of polydisperse dextrans was used to

    characterize the retention properties of the templatedcarbon films. Dextrans having average molecular

    weights of 2 106, 1.7 105, 7.0 104, 3.9 104,

    9.9 103 gmol1 (Pharmacia) were obtained from

    Aldrich Chemical Company. The dextrans were added

    in equal mass ratios to create a 10 g l1 solution in a

    phosphate buffer prepared from monobasic (NaHPO4)

    and dibasic (NaH2PO4) sodium phosphate to achieve

    a solution pH of 7.40.3. When applicable, 1 wt.% of

    methanol was added to prevent bacterial contamina-

    tion.

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    5/14

    M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186 177

    Retentates and permeates were analyzed using size

    exclusion chromatography (SEC) with a model 1050

    HP chromatograph (Hewlett Packard) and a Waters

    410 differential refractometer. SEC analysis wasperformed using a 7.5-mm i.d. 30-cm TSK G4000

    SW column (Tosoh Corp.), thermostated at 30 C,

    with 0.1 M sodium sulfate used as the mobile phase

    at a flow rate of 0.6 ml min1. Samples were injected

    with an autosampler without additional dilution using

    a 100-l sample loop. The above mentioned dex-

    trans were injected individually at a concentration of

    1mgml1, and retention volume data were used to

    construct a molecular weight calibration curve. Data

    was processed using Waters Millennium 32 HPLC

    software.

    Membrane separation was carried out using a

    dead end filtration setup. A 50 ml volume of dextran

    solution was loaded above the carbon membrane in

    a 30 mm diameter Amicon stirred-cell. An attached

    Ar line supplied gas pressure to the solution above

    the membrane. The permeate was collected in 30 ml

    vials, with the total permeate mass weighed contin-

    uously by placing the vial directly on a laboratory

    balance. From this time series data mass flow through

    the membrane was computed. A series of permeate

    and retentate samples were collected over time for

    off-line dextran analysis.

    2.5. Batch depletion dextran adsorption

    Granular carbon membrane samples were prepared

    using the identical procedure described above for

    making thin carbon films, except that the solution

    was poured into a quartz boat and pyrolyzed as a

    bulk solution. These samples were analyzed for total

    dextran adsorption capacity using a batch depletion

    method. Each carbon was ground to 60/140 mesh

    and suspended in a water solution. After settling, theremaining solution was decanted to remove fine parti-

    cles. The process was repeated until all fine particles

    were removed. A mass of 0.5 g of carbon was sealed

    with 30 ml of the polydisperse dextran solution in

    a Pyrex vial as described above. Sealed vials were

    vigorously agitated for 32 h, after which the solutions

    were passed through a 1m filter and analyzed using

    SEC as described above. The average concentration

    of dextran remaining in the solution above the carbon

    was integrated over the molecular weight distribu-

    tion, with the results compared to the corresponding

    integral for the original solution.

    3. Results and discussion

    Table 1 lists the initial support weight, the weight

    after spray coating (wet), and the weight after car-

    bonization (dry) for each membrane studied in this

    work. All samples except M8000-3-0.022 were given

    a single coat and thermal treatment. This particular

    sample was fabricated in two coating/carbonization

    cycles. During the carbonization process, the evolu-

    tion of gases reduced the polymer mass on the support

    leaving behind the solid carbon film. For poly(furfuryl

    alcohol) carbonization at 600 C, the carbon yield pro-duced in this way is approximately 2530% for sim-

    ilar thin films [14]. The PEG template is expected to

    produce very little residual carbon as seen in the syn-

    thesis of mesoporous catalyst supports (

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    6/14

    178 M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186

    Fig. 3. Water permeance through carbon membranes as a functionof template molecular weight.

    the entire template being emitted as gaseous prod-

    ucts. This suggests that volatilization of the template

    does not cause any significant loss of poly(furfuryl

    alcohol) prior to carbonization.

    It is noteworthy that a similar trend is observed for

    the water permeability (Fig. 3) plotted as a function of

    Fig. 4. Scanning electron micrograph of a porous carbon ultrafiltration membrane (M8000-1-0.021) cross-section showing external carbon

    layer and stainless steel support.

    template size although there is considerable variabil-

    ity between samples formed from the same template

    size. Generally, water permeability decreases with

    decreasing template size and this may correspond toa decrease in membrane pore size and porosity as dis-

    cussed below. In Table 1, the values for the hydraulic

    permeabilities range from 1014 to 1011 m s1 Pa1,

    which are considerably lower than commercial ul-

    trafiltration membranes (typical values are around

    1010 m s1 Pa1. These reduced permeabilities are

    attributable to both the much greater thickness of the

    membranes and the decrease in the film porosity that

    occurs with decreasing template size. These factors

    are discussed in detail below.

    3.1. Scanning electron microscopy

    Fig. 4 is a scanning electron micrograph of a

    cross-section of the external porous carbon layer and

    the stainless steel support for a membrane synthesized

    using a 50% mixture of PFA: 8000 g mol1 PEG at

    600 C. The demarcation between the stainless steel

    support and external carbon layer is clearly visible,

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    7/14

    M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186 179

    and there are also localized along with localized sep-

    arations or cracks between the two layers. These film

    separations apparently have a negligible impact on

    coating integrity and mechanical stability, as therewas no observed carbon mass loss during testing or

    property variations after repeated testing. The com-

    bination of the diamond saw cutting and surface

    polishing of the SEM sample has a smoothing affect

    on the morphology of the membrane cross-sections.

    The actual porosity of the steel support is roughly

    0.6 as reported by the manufacturer. The carbon layer

    is 12- to 15-m thick and relatively uniform over

    the membrane surface, consistent with observations

    of similarly prepared carbon films for gas separation

    [14]. The calculated thickness range based on the

    mass of carbon deposited (Table 1) and assuming

    a porous carbon density of 1.6 g cm3 is 915m,

    which agrees well with this measurement.

    Fig. 5 is a 500 nm scale micrograph of the mi-

    crostructure of the PFA/PEG carbon taken from

    the material external to the support surface. This

    cross-section was prepared by fracturing the carbon

    surface while attached to the support using a scalpel.

    Fig. 5. Scanning electron micrograph of the pore structure of the carbon ultrafiltration membrane showing grain boundaries between

    nanoporous carbon domains (M8000-1-0.021).

    The microstructure suggests pores with diameters in

    the nanometer range. As expected from the inferred

    hypothesis for pore formation, these pores are in the

    form of fissures and voids between purely nanoporouscarbon domains. Porous carbon was also observed to

    be present on the opposite side of the support (the

    side opposite from the initial deposition) and may

    fill some portion of the macroporosity of the sup-

    port. During pyrolysis the polymer blend drops in

    viscosity prior to carbonization, making deeper pore

    penetration more facile.

    3.2. Porosity characterization by gas permeation

    Fig. 6(a) is a plot of the flux versus pressure for

    several gases through an uncoated stainless steel

    0.2m support. The data show both a high rate of

    permeation and low gas selectivity. A nanoporous

    carbon film pyrolized in the absence of polymer tem-

    plate shows a linear relationship between the flux and

    applied pressure for He, O2, N2 and Ar (Fig. 6(b)).

    This permeation behavior is typical of PFA derived

    carbon membranes [18]. Fig. 7 is characteristic of gas

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    8/14

    180 M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186

    Fig. 6. Flux as a function of driving force pressure for: (a) uncoated stainless steel support; (b) 0% PEG/PFA carbon membrane carbonized

    at 600 C; (c) 50% 8000 amu PEG/PFA templated membrane carbonized at 600 C.

    permeation through a templated carbon membrane

    (50% PFA/8000 g mol

    1 PEG) showing a quadraticdependence of the flux versus pressure and Knudsen

    gas selectivity at low pressure. This behavior is

    Fig. 7. Ultrafiltration membrane (M3400-1-0.019) gas permeability

    as a function of trans-membrane pressure for mean pore size

    characterization.

    observed on all PEG/PFA carbon membranes regard-

    less of coating deposition mass or template size emplo-yed. Hence, the permeation results cannot necessarily

    be attributed to partially coated or partially formed

    nanoporous films such as that shown in Fig. 6(b).

    For a membrane with a mean pore size between

    100 and 1 nm, the pore dimension can be estimated

    by examining data for the membrane gas perme-

    ability versus trans-membrane pressure. The linear

    relationship yields an intercept equal to the Knud-

    sen permeability with slope inversely proportional to

    the gas viscosity (Fig. 7). The gas flux through the

    membrane is written as

    J=

    K0 +B0

    p

    p

    l(1)

    The structure factor B0 is proportional to the square

    of the mean pore size

    B0 =

    2d2m (2)

    with = 2.5 for consolidated media [20]. The

    Knudsen permeability K0 can be written as

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    9/14

    M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186 181

    Table 2

    Mean pore sizes of membranes measured by gas permeation

    dm (m)

    PEG1500 PEG3400 PEG8000

    He 9.92 109 1.19 108

    N2 1.03 108 1.28 108 2.83 108

    Ar 1.37 108 6.81 109 2.59 108

    O2 9.15 109 2.39 108 2.21 108

    CO2 1.43 108 1.22 108 2.87 108

    SF6 1.13 108 1.10 108

    Mean 1.15 108 1.31 108 2.63 108

    S.D. 2.10 109 5.69 109 3.05 109

    K0 =4

    32 dm8RTM (3)

    where is 0.8 for consolidated media [20]. The gas

    permeation data can be used to evaluate both B0 and

    K0, with the ratio of these two parameters used to

    evaluate the mean pore size of the membrane, dmwithout any assumptions regarding the membrane

    porosity, or tortuosity factor, . Here, we do assume

    that the pore size of the carbon film is significantly

    less than that of the support (i.e.

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    10/14

    182 M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186

    Fig. 8. Observed (dotted line) and actual (solid line) partial rejection coefficients vs. dextran molecular weight for carbon membranes

    synthesized from poly(ethylene glycol) template: (a) 1000 MW PEG; (b) 1500 MW PEG; (c) 2000 MW PEG; (d) 3400 MW PEG; (e)

    8000 MW PEG.

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    11/14

    M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186 183

    as a function of dextran molecular weight for mem-

    branes synthesized using the five different template

    PEG sizes. The results for 1000 MW PEG shown

    in Fig. 8(a) demonstrate a lack of dextran retentionwith respect to molecular weight indicating that large,

    non-selective pores exist across the film. The polar-

    ization correction reveals that the broad, non-selective

    behavior of the membrane cannot be attributed to po-

    larization, but rather that the membrane is defective.

    As the template size tends toward zero, the carbon

    membrane behaves as an untemplated film with a sig-

    nificant tendency towards cracking. It has been well

    established that furfuryl alcohol resin-based mem-

    branes require several coatings before a crack-free

    layer can be constructed [12,14]. The curves for

    1500 (Fig. 8(b)) and 2000 PEG templates (Fig. 8(c))

    show an increase in membrane selectivity. However,

    a tendency towards film cracking is retained and

    membrane reproducibility is affected. Fig. 8(c) in

    particular shows a lack of reproducibility for four

    membranes synthesized at the same carbon deposition

    mass. Membrane M2000-4-0.019 (Fig. 8(c)) shows a

    significantly enhanced retention compared to those in

    Fig. 8(a), which may represent the properties of the

    film in the absence of any significant cracking. Such

    retention could not be attained using a 1000 MW pre-

    cursor. The templating of nanoporous carbon usingPEG has been observed to yield accessible meso-

    porosity only at molecular weights of 2000 g mol1

    and higher for granular materials [15], although

    this study represents the most comprehensive work

    undertaken to determine the exact threshold.

    Problems of film cracking and reproducibility

    diminish for higher molecular weight templates as

    transport through these films becomes dominated by

    that which takes place through the porosity created via

    the templates rather than through adventitious cracks

    and fissures. Fig. 8(d) and (e) show reproduciblebehavior for 3400 and 8000 MW PEG templated

    membranes, respectively. Retention properties tend to

    increase from those seen in Fig. 8(b) and (c) and are

    apparently independent of carbon thickness and water

    permeance as expected. Membrane M8000-3-0.022 is

    the result of two successive coating and carbonization

    cycles and shows an increase in retention behavior

    likely from a reduction in pore size after the initial

    coat. Here, the pore structure of the first coat may be-

    come impregnated with a second coating of precursor

    Fig. 9. Dextran R90 value and calculated pore size as a function

    of template molecular weight of carbon membrane.

    material that modifies the retention characteristics

    as shown. This procedure, which is analogous to a

    slip casting design, may in fact provide a route to

    the synthesis of ultra-thin nanofiltration membranes.

    With successive layer deposition, the retentive prop-

    erties decrease with only a marginal increase in film

    thickness and deposition mass.

    In Fig. 9, the molecular weight of dextran corre-sponding to 90% rejection is plotted as a function of

    the molecular weight of the PEG template. This R90value is obtained by interpolation or extrapolation of

    the data in Fig. 8(a)(e) as needed. The extrapolated

    value for the 1000 g mol1 template determined from

    Fig. 8(a) is actually an underestimate of this molecular

    weight. As observed above, the analysis suggests that

    these films possess large cracks and defects that are

    non-retentive. The trend shows an increase in dextranR90 with increasing PEG molecular weight starting

    from the 2000 MW PEG template size. Below thiscutoff, the selectivity decreases sharply with decreas-

    ing template size suggesting the formation of cracks

    and defects. Fig. 9 also plots the calculated pore size

    based upon a simple partitioning model assuming a

    cylindrical pore structure and using the Stokes radius

    of the dextran molecule for the corresponding R90molecular weight [1]

    Rc = 1

    1rdextran

    rpore

    2(6)

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    12/14

    184 M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186

    Pore sizes estimated in this way for the membranes

    formed from the larger PEG templates are generally

    smaller than those calculated from gas permeability

    data. For example, the values of 15 and 12.4 nm cor-responding to 8000 and 3400 molecular weight PEG

    compare to 26.3 6.1 and 13.1 11.4 nm calculated

    from gas permeation, respectively. In contrast, the

    pore size calculated from the dextran data for the

    1500 PEG membrane was 28.2 nm, which is signifi-

    cantly greater than the 11.54.2 nm determined from

    gas permeation. This large discrepancy is due to the

    greater effect of the pore defects on dextran transport

    than on the gas permeation results.

    The water permeability can also be used to provide

    an estimate of the average pore size or porosity using

    the HagenPoiseuille equation. Here, the membrane

    porosity/tortuosity ratio, /, can be expressed as a

    function of the pore size and water permeability Lpassuming a cylindrical pore structure [1]

    =

    8lLp

    (rpore)2(7)

    Using pore sizes from dextran rejection data (Eq. (6)),

    application of Eq. (7) predicts that the ratio of porosity

    to tortuosity decreases with decreasing template size.

    Average values are 0.024, 0.005, 0.001 and 0.001 for

    8000, 3400, 2000, and 1000 g mol1 template. Thisdecrease in / is consistent with the greater carbon

    yield, and thus, lower porosity, of the membranes

    formed using the smaller PEG templates.

    Results of the batch depletion of dextran over gran-

    ular templated carbons yield relative values of the

    accessible pore volumes of the various carbons. This

    allows for a comparison between properties resulting

    from the templating process and those resulting from

    pores developed as a result of film cracking. Fig. 10

    is a plot of relative dextran amount adsorbed versus

    PEG molecular weight used during carbon synthesis.The results show negligible uptake at values of 2000

    MW PEG and below with substantial increases in

    accessible area at larger template molecular weights

    (sizes). The data suggest a maximum in adsorption

    uptake near 8000 MW PEG. At 18,500 MW PEG, the

    extent of adsorption falls to values below that of the

    3400 MW PEG. The initial portion of the curve agrees

    well with previous gas adsorption characterization,

    where it has also been noted that accessibility of the

    mesopore distribution becomes hampered for lower

    Fig. 10. Dextran uptake from batch depletion on granular carbons

    synthesized using templates of 30018,500 g mol1 of PEG.

    template molecular weights [15]. It is this lack of ac-

    cessibility that may be responsible for the trapping

    of template material within the carbonizing system

    and hence the increase in carbon yield at low template

    molecular weights (Fig. 2). This retention of material

    would explain the observed decrease in membrane

    porosity. In Fig. 10, the subsequent decrease in capac-

    ity with 18,500 PEG may be the effect of a decrease inthe rate of pyrolysis of this larger polymeric template.

    The subsequent decrease in the rate of evolution of the

    pyrolysis gases could also lead to a retention of tem-

    plate carbon within the forming solid, as in the case

    of the diffusion limitations associated with smaller

    template sizes.

    The internal pore volume of membranes made from

    the smaller templates is largely inaccessible, leading

    to a small amount of dextran uptake. As the template

    size increases, the accessibility of the pore volume

    increases, but at the expense of an increasing aver-age pore radius, which ultimately reduces the internal

    surface area.

    The emerging picture of the actual templating

    process from the above observations is one that is

    kinetically driven by the diffusion of template decom-

    position products from the forming solid. At small

    template pore sizes (

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    13/14

    M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186 185

    resistance is exceedingly high. These diffusion lim-

    itations increase carbon yield and decrease carbon

    porosity as observed. The lack of mesopore intercon-

    nection also means that the membrane behaves as apurely nanoporous thin film, and hence is subject to

    cracking and the formation of defects after a single

    coating. As the template size increases, this diffusion

    length scale increases and membrane properties be-

    come a function of the templated carbon itself with a

    subsequent improvement in retention properties and

    reproducibility.

    The results presented here underscore several

    potential improvements that can be made through

    modification of the synthesis procedure. The most

    notable of these is improvement in water perme-

    ability by decreasing carbon film thickness. The

    spray deposition on macroporous support was readily

    adapted from an analogous method of gas-separation

    membrane fabrication [15], although this may not be

    the optimal system for liquid separation applications.

    The results of microscopy and water permeation mea-

    surements suggest that decreasing the carbon deposi-

    tion mass while maintaining coating uniformity can

    accomplish this goal. Additionally, this work presents

    indirect evidence that lower molecular weight tem-

    plates have a discernable affect on pore structure and

    may not be entirely flashed off during carbonizationas originally proposed [17]. This warrants a more

    complete investigation of the templating process with

    a particular emphasis on the composition and molec-

    ular weight dependence of the template as well as the

    pyrolysis ramp rate during carbonization.

    4. Conclusions

    A novel method for producing ultra and nanofil-

    tration carbon membranes is presented using spraydeposition on macroporous stainless steel and pyrol-

    ysis of poly(furfuryl alcohol)/poly(ethylene glycol)

    mixtures. Membranes characterized by SEM show

    12- to 15-m thick carbon films. Pores in the ultrafil-

    tration range clearly vary with the average molecular

    weight of the template for PEG with molecular weight

    of 2000 g mol1 and above. Below this threshold, the

    film behaves similar to an untemplated carbon mem-

    brane with crack formation a significant factor for the

    single coated membranes. Transport of a polydisperse

    dextran solution was used to quantify the retention

    properties of the membranes. Partial rejection coef-

    ficients for 90% of the dextran mixture were shown

    to vary with template molecular weight: 90% cutoffsof 2 104, 3.5 104, 6 104 gmol1 of dextran

    were measured for 2000, 3400, and 8000 g mol1 of

    template employed, respectively. Although generally

    of less accuracy, the mean pore sizes as measured

    from gas permeation data also demonstrate a similar

    decrease in value with decreasing template size.

    Acknowledgements

    Support for this research was provided by the

    Department of Energy Office of Basic Energy Science

    and the DuPont Co. Michael Strano is grateful for fi-

    nancial support in the form of a Presidential Graduate

    Fellowship from the University of Delaware.

    References

    [1] L.J. Zeman, A.L. Zydney, Microfiltration and Ultrafiltration:

    Principles and Applications, Marcel Dekker, New York, 1996.

    [2] A.S. Michaels, High flow membrane, US Patent #3615024

    (1971).

    [3] T.A. Tweddle, W.L. Thayer, O. Kutowy, S. Sourirajan, Methodof gelling cast, polysulfone memrbrane, US Patent #4451424

    (1984).

    [4] E. Schindler, F. Maier, Manufacture of porous carbon

    membranes, US Patent #4919860 (1990).

    [5] W.M. Clark, A. Bansal, M. Sontakke, Y.H. Ma, Protein

    adsorption and fouling in ceramic ultrafiltration membranes,

    J. Membr. Sci. 55 (1991) 2138.

    [6] H.P. Hsieh, R.R. Bhave, H.L. Fleming, Microporous alumina

    membranes, J. Membr. Sci. 39 (1988) 221241.

    [7] J.C.S. Wu, L.C. Cheng, An improved synthesis of ultrafil-

    tration zirconia membranes via the solgel route using

    alkoxide precursor, J. Membr. Sci. 167 (2) (2000) 253261.

    [8] H.C. Foley, Carbogenic molecular-sieves synthesis, proper-ties and applications, Micropor. Mater. 4 (6) (1995) 407433.

    [9] H.C. Foley, Nanoporous carbons and related materials for

    small molecule separations, Abstr. Papers Am. Chem. Soc.

    211 (1996) 2.

    [10] M. Acharya, M.S. Strano, J. Mathews, S.J.L. Billinge, et al.,

    Simulation of nanoporous carbons: a chemically constrained

    structure, Philos. Mag. B 79 (10) (1999) 14991518.

    [11] R.K. Mariwala, H.C. Foley, Calculation of micropore sizes

    in carbogenic materials from the methyl-chloride adsorption-

    isotherm, Ind. Eng. Chem. Res. 33 (10) (1994) 23142321.

    [12] M. Acharya, B.A. Raich, H.C. Foley, M.P. Harold, J.J. Lerou,

    Metal-supported carbogenic molecular sieve membranes:

  • 7/27/2019 Ultrafiltration Membrane Synthesis by Nanoscale Templating

    14/14

    186 M.S. Strano et al. / Journal of Membrane Science 198 (2002) 173186

    synthesis and applications, Ind. Eng. Chem. Res. 36 (8) (1997)

    29242930.

    [13] M. Acharya, H.C. Foley, Spray-coating of nanoporous carbon

    membranes for air separation, J. Membr. Sci. 161 (1999) 15.

    [14] M.B. Shiflett, H.C. Foley, Ultrasonic deposition of highselectivity nanoporous carbon membranes, Science 285 (17)

    (1999) 19021905.

    [15] D.S. Lafyatis, J. Tung, H.C. Foley, Poly(furfuryl alcohol)-

    derived carbon molecular-sieves dependence of

    adsorptive properties on carbonization temperature, time and

    poly(ethylene glycol) additives, Ind. Eng. Chem. Res. 30 (5)

    (1991) 865873.

    [16] M.S. Kane, L.C. Kao, R.K. Mariwala, D.F. Hilscher, H.C.

    Foley, Effect of porosity of carbogenic molecular sieve

    catalysts on ethylbenzene oxidative dehydrogenation, Ind.

    Eng. Chem. Res. 35 (10) (1996) 33193331.

    [17] M.G. Stevens, H.C. Foley, Alkali metals on nanoporous

    carbon: new solid-base catalysts, Chem. Commun. (6) (1997)

    519520.

    [18] M.S. Strano, H.C. Foley, Deconvolution of permeance through

    supported nanoporous membranes, AIChE J. 46 (3) (2000)

    651658.

    [19] M. Stevens, Cesium/nanoporous carbon composite materials:

    synthesis, characterization, and base catalysis, Ph.D. Thesis,University of Delaware, Newark, 1999 (Chapter 2).

    [20] P.C. Carman, Flow of Gases through Porous Media, Academic

    Press, London, 1956.

    [21] R. Nobrega, H. Balmann, P. Aimar, V. Sanchez, Transfer of

    dextran through ultrafiltration membranes: a study of rejection

    data analysed by gel permeation chromatography, J. Membr.

    Sci. 45 (1989) 1736.

    [22] K.A. Granath, Solution properties of branched dextrans, J.

    Colloid Interf. Sci. 13 (1958) 308.

    [23] K.A. Smith, C.K. Colton, E.W. Merrill, L.B. Evans,

    Convective transport in a batch dialyzer: determination of

    the true membrane permeability from a single measurement,

    AIChE Symp. Ser. 64 (1968) 45.