Aiello Agbogbo y Hotzapple BTech 2006

download Aiello Agbogbo y Hotzapple BTech 2006

of 10

Transcript of Aiello Agbogbo y Hotzapple BTech 2006

  • 8/8/2019 Aiello Agbogbo y Hotzapple BTech 2006

    1/10

    Conversion of municipal solid waste to carboxylic acidsusing a mixed culture of mesophilic microorganisms

    Cateryna Aiello-Mazzarri a, Frank K. Agbogbo b, Mark T. Holtzapple b,*

    a Universidad del Zulia, Facultad de Ingeniera, Coordinacion de Ciclo Basico, Departamento de Qumica, Maracaibo, Venezuelab Texas A&M University, Department of Chemical Engineering, College Station, TX 77840-3122, USA

    Received 22 October 2004; received in revised form 10 February 2005; accepted 11 February 2005

    Available online 12 April 2005

    Abstract

    Waste biomass was anaerobically converted to carboxylate salts by using a mixed culture of acid-forming microorganisms. Muni-

    cipal solid waste (MSW) was the energy source (carbohydrates) and sewage sludge (SS) was the nutrient source (minerals, metals,

    and vitamins). Four fermentors were arranged in series and solids and liquids were transferred countercurrently in opposite direc-

    tions, which allows both high conversions and high product concentrations. Fresh biomass was added to Fermentor 1 (highest car-

    boxylic acid concentration) and fresh media was added to Fermentor 4 (most digested biomass). All fermentations were performed

    at 40 C. Calcium carbonate was added to the fermentors to neutralize the acids to their corresponding carboxylate salts. Iodoform

    was used to inhibit methane production and urea was added as a nitrogen source. Product concentrations were up to 25 g/L, with

    productivities up to 1.4 g total acid/(L liquid d). Mass balances with closure between 93% and 105% were obtained for all systems.

    Continuum particle distribution modeling (CPDM) was applied to correlate batch fermentation data to countercurrent fermentation

    data and predict product concentration over a wide range of solids loading rates and residence times. CPDM for lime-treated MSW/

    SS fermentation system predicted the experimental total acid concentration and conversion within 4% and 16% respectively.

    2005 Published by Elsevier Ltd.

    Keywords: MixAlco process; Municipal solid waste; Sewage sludge; Fermentation; Carboxylic acids; Mixed acids

    1. Introduction

    Many energy experts are concerned about the deple-

    tion of petroleum, the main source of liquid fuels and

    many chemicals. Oil prices are volatile and supplies

    are unstable. Oil combustion releases carbon dioxide,

    the most important greenhouse gas implicated in globalwarming (Sterzinger, 1995; Klass, 1998). In contrast,

    biomass is a source of liquid fuels that does not result

    in a net increase of carbon dioxide in the atmosphere

    (Holtzapple et al., 1997; Hileman, 1999) because bio-

    mass growth removes from the atmosphere the same

    amount of carbon dioxide that biomass combustion

    generates (Bungay, 1981; Sterzinger, 1995).

    Another environmental problem is the generation

    and accumulation of the following wastes: Municipal

    solid waste (MSW)in the United States (US), ap-

    proximately 220 million tons of MSW were generated

    in 1998, four million tons more than in 1997 (EPA,1998). Municipal solid waste contains about 65% biode-

    gradable components, such as paper, food scraps, and

    yard waste (EPA, 1998). For the purpose of this

    research, we will refer to MSW as the biodegradable

    organic fraction, which excludes glass, metal, plastic,

    ceramic, rock, or soil also found in MSW. Sewage

    sludge (SS)SS is the residual solids from conventional

    aerobic or anaerobic sewage treatment. The Environ-

    mental Protection Agency (EPA, 1999) estimated that

    0960-8524/$ - see front matter 2005 Published by Elsevier Ltd.

    doi:10.1016/j.biortech.2005.02.020

    * Corresponding author. Tel.: +1 979 845 9708; fax: +1 979 845 6446.

    E-mail addresses: [email protected] (C. Aiello-Mazzarri), m-holtzapple@

    tamu.edu (M.T. Holtzapple).

    Bioresource Technology 97 (2006) 4756

    mailto:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]
  • 8/8/2019 Aiello Agbogbo y Hotzapple BTech 2006

    2/10

  • 8/8/2019 Aiello Agbogbo y Hotzapple BTech 2006

    3/10

    apple et al., 1992). The components of MSW were col-

    lected (Bryan, College Station, TX), sun-dried, ground

    and passed through a 10-mm screen in a hammer mill

    (Forest Science Research Laboratory, Texas A&M Uni-

    versity). Once combined, the MSW was ground and

    passed through a 6-mm screen to ensure a uniform mix-

    ture. Fats and oils were not added to the MSW to pre-vent spoiling during storage.

    2.2.2. Sewage sludge

    Aerobically treated sewage sludge was obtained from

    Bryan Wastewater Treatment Plant Number 3 (Bryan,

    TX). The sewage had undergone activated sludge treat-

    ment in which the incoming sewage stream was sent to

    an aeration basin for approximately 15 d. After leav-

    ing the aeration basin, the sludge was digested aero-

    bically for approximately 40 d. Upon removal from

    the digestor, the sludge was coagulated, dried for 10 d,

    and ground in a hammer mill fitted with a 3-mm screen.

    2.3. Lime treatment

    The components of lignocellulose are complex and

    include solubles, cellulose, hemicellulose, lignin, and

    ash. As biomass digests, the remaining fraction is less

    reactive (Gizjen et al., 1988), which is logical considering

    that biomass is heterogeneous; some components are

    highly digestible whereas others are almost inert. To

    enhance digestibility, lime treatment was performed at

    100 C for 1 h with loadings of 0.1 g Ca(OH)2/g dry bio-

    mass and 10 mL of distilled water/g dry biomass. After

    treatment, CO2 was bubbled through the biomass slurryto neutralize the lime and the slurry was dried at 105 C.

    Detailed procedures of the lime treatment are available

    in Aiello-Mazzarri (2002).

    2.4. Media and nutrients

    The fermentation media consisted of deoxygenated

    distilled water, 0.28 g sodium sulfide/L distilled water,

    and 0.28 g cysteine hydrochloride/L distilled water.

    Dry nutrients were added to the fermentations. The

    dry nutrient mixture used in all the experiments corre-

    sponds to the modified Caldwell and Bryant medium

    (Caldwell and Bryant, 1966).

    2.5. Inocula

    Rumen fluid was used as the main inoculum. Rumen

    contents were removed from a fistulated steer (Univer-

    sity Nutrition and Field Laboratory, Texas A&M Uni-

    versity), filtered through four layers of cheesecloth into

    1-L propylene centrifuge bottles, transported to the lab-

    oratory, and inoculated (5% v/v) into each fermentor

    within 1 h after collection to ensure microorganism via-

    bility. Other sources of microorganisms included swamp

    material from Bee Creek Park (College Station, TX),

    and compost material from domestic and commercial

    piles. The compost samples were taken from the middle

    of the pile to ensure a large quantity of anaerobic micro-

    organisms. To minimize exposure to oxygen, the swamp

    material and compost were collected into bottles filled

    with deoxygenated distilled water.

    2.6. Methanogens inhibitor

    Iodoform (CHI3) was used as a methanogen inhibitor

    in all fermentations. An iodoform solution (20 g CHI3/L

    ethanol) was added individually to each reactor contin-

    uously throughout the fermentations. Iodoform is light

    and air sensitive, so the solution was kept in amber-

    colored glass bottles and special care was taken to

    replace the cap immediately after use.

    2.7. Batch experiments

    Batch experiments were performed to determine

    experimental reproducibility, and to collect the neces-

    sary data for continuum particle distribution modeling

    (CPDM). These data were used to develop a model rate

    equation for the countercurrent fermentation. The pro-

    cedure involved operating five different fermentors with

    different initial substrate concentrations (20, 40, 70,

    100 g dry substrate/L liquid). An additional fermentor

    with high substrate concentration (100 g dry substrate/

    L liquid) and high carboxylic acid concentration

    (approximately 20 g carboxylic acids/L liquid) was eval-

    uated. Samples were taken by removing the fermentorfrom the incubator, collecting the gas, removing the

    stopper, purging with nitrogen, measuring pH, centri-

    fuging for 10 min, collecting a liquid sample under nitro-

    gen purge, resealing the fermentor, and placing it back

    in the roller apparatus. For the CPDM experiments,

    the resulting carboxylic acid concentrations were con-

    verted into acetic acid equivalents to be used in the

    modeling equations developed by Loescher (1996).

    The batch carboxylic acid concentrations can be con-

    verted to acetic acid equivalents (a)

    a mol=L acetic mol=L

    1:75 propionic mol=L

    2:5 butyric mol=L

    3:25 valeric mol=L

    4:0 caprioc mol=L

    4:75 heptanoic mol=L 1

    On mass basis, the acetic acid equivalent can be ex-

    pressed as

    Ae g=L 60:05 g=mol a mol=L 2

    C. Aiello-Mazzarri et al. / Bioresource Technology 97 (2006) 4756 49

  • 8/8/2019 Aiello Agbogbo y Hotzapple BTech 2006

    4/10

    2.8. Countercurrent fermentations

    In countercurrent operation, liquid and solids flow in

    opposite directions in four-fermentor trains. In the labo-

    ratory scale, the fermentors operate in a semi-continuous

    manner, whereas on an industrial scale, the fermentation

    would be continuous. Countercurrent fermentations wereinitiated as batch cultures under anaerobic conditions by

    adding the substrates, calcium carbonate, urea, nutrients,

    and inocula to deoxygenated water media in each fer-

    mentor. The experiments were conducted as batch fer-

    mentations until the culture was established (710 d).

    Countercurrent operation was initiated with the transfer

    of liquid and solids occurring every 1, 2, or 3 d. Counter-

    current fermentations were conducted at varying liquid

    residence times (LRT) and volatile solid loading rate

    (VSLR). The operating parameters for each fermentation

    with lime-treated MSW/SS are shown in Table 1.

    2.9. Reaction conditions

    The fermentations were performed under anaerobic

    conditions at 40 C. Every 3 d, after each liquid/solid

    transfer, 2.0 g calcium carbonate was added to each fer-

    mentor to neutralize carboxylic acids. To maintain

    anaerobic conditions, nitrogen from a high-pressure

    liquid-nitrogen cylinder (Praxair, Bryan, TX) was flushed

    whenever the fermentors were open to the atmosphere.

    The solid and liquid transfer procedures are detailed in

    Aiello-Mazzarri (2002).

    2.10. Analytical methods

    The fermentor broth was analyzed by gas chromatog-

    raphy to measure the concentration of carboxylic acids.

    The broth was mixed with equal parts of an internal

    standard (4-methyl-n-valeric acid) and 3-M H3PO4.

    The analysis was performed using an Agilent 6890 series

    gas chromatograph (Agilent Technologies, Palo Alto,

    CA) equipped with a flame ionization detector (FID)

    and a 7683 series injector. A 30-m fused-silica capillary

    column (J&W Scientific, Model # 123-3232 CX, AgilentTechnologies, CA) was used. The column head pressure

    was maintained between 90 and 103 kPa (1315 psig).

    At every sample injection, the gas chromatograph tem-

    perature program allowed the temperature to rise from

    50 C to 200 C at a 20 C/min rate. The temperature

    was subsequently held at 200 C for 10 min. Helium

    was used as carrier gas, and the total run time per sam-

    ple was 17 min. The fermentation broth consists of a

    mixture of carboxylate salts and carboxylic acids. This

    analytical procedure converts all salts to their corre-

    sponding acids, allowing product concentrations to be

    reported as g carboxylic acid/L.

    Gases produced during fermentation were accumu-

    lated within the reactor. Every sampling day, the volume

    of gas produced since the last transfer session was mea-

    sured. The volume was measured by displacing water in

    an inverted glass graduated cylinder apparatus that was

    filled with 30% CaCl2 solution. The CaCl2 minimized

    microbial growth in the water tank, and reduced water

    evaporation. The CaCl2 solution had an acidic pH

    (5.6), which prevented CO2 adsorption. To check meth-

    anogen inhibition, every 2 or 3 d the gas from the

    fermentation was analyzed for CH4 by gas chromatog-

    raphy. A 5-mL sample was taken through the reactor

    septum and analyzed using an Agilent 6890 series gaschromatograph. The chromatograph was equipped with

    a thermal conductivity detector (TCD). A 4.6-m stain-

    less steel packed column with 2.1-mm ID (60/80 Carbo-

    xen 1000, Supelco 1-2390 U, Agilent Technologies,

    CA) was used. Samples were injected manually. The

    inlet temperature was fixed at 230 C, and the detector

    temperature was set at 250 C. The oven temperature

    was maintained at 225 C for 5 min. Helium was used

    as carrier gas. The total elution time for a sample was

    5 min.

    Volatile solids in the initial substrates and solid fer-

    mentation residues were determined by first drying the

    material at 105 C and then ashing the material at

    550 C for at least 3 h. Volatile solid determination

    in the liquid fermentation broth followed the same

    heating procedure above; except prior to drying, the

    liquid was mixed with lime to ensure that the carbox-

    ylic acids would not volatize and alter the measure-

    ment.

    2.11. Mass balance

    For all the countercurrent fermentation experiments,

    a complete mass balance was obtained on the entire

    Table 1

    Operating parameters for lime-treated MSW/SS countercurrent

    fermentations

    Fermentation trains A B C D E

    LRT (d) 20.7 19.0 22.7 20.7 17.0

    VSLR (g VS/(L liquid in

    all fermentors d))

    3.0 4.0 8.1 3.1 6.4

    VS feed at each

    transfer (g VS)

    11.6 19.3 24.1 11.6 15.4

    Solid feed at each

    transfer (g dry)

    14.4 24.0 30.0 14.4 19.2

    Liquid feed to F4 at

    each transfer (L)

    0.15 0.20 0.10 0.15 0.10

    Frequency of transfer Every 3 d

    Iodoform addition rate

    (mg iodoform added/L

    liquid fed to F4)

    5.33 8 8 5.33 8

    Nutrients addition rate

    (g dry nutrients added/L

    liquid fed to F4)

    1.33 1.00 2.00 1.33 2.00

    Urea addition rate

    (g urea added/L

    liquid feed to F4)

    1.0 0.75 1.5 1.0 1.5

    50 C. Aiello-Mazzarri et al. / Bioresource Technology 97 (2006) 4756

  • 8/8/2019 Aiello Agbogbo y Hotzapple BTech 2006

    5/10

    train during a steady-state period. Closure is defined as

    follows:

    closure mass out

    mass in water of hydrolysis3

    To calculate the water of hydrolysis, it was assumed that

    the biomass could be represented as cellulose, which has

    a monomer weight of 162 g/mol. When cellulose is

    hydrolyzed, it gains one molecule of water per mono-

    mer; therefore, the water of hydrolysis is calculated as

    water of hydrolysis VS digested 18162 5

    2.12. Operational parameters

    The liquid residence time determines how long the

    liquid remains in the system, which affects the final prod-

    uct concentration. Long liquid residence times allow

    high product concentrations whereas shorter liquid

    residence times allow lower product concentrations

    (Holtzapple et al., 1999). Liquid residence time is calcu-

    lated as

    liquid residence time LRT TLVQ

    6

    where, Q = flowrate of liquid out of the fermentor set

    (L/d), TLV = total liquid volume.

    TLV X

    i

    Ki w Fi 7

    where, Ki average wet mass of solid cake in Fermentori g; w = average liquid fraction of solid cake inFermentor i (L liquid/g wet cake); Fi averagevolume of free liquid in Fermentor i L.

    The volatile solids loading rate is calculated as

    volatile solids loading rate VSLR VS fed=dTLV

    8

    At a low VSLR, the solid residence time increases,

    allowing for more complete digestion.

    Biomass is composed of volatile solids (VS) and ash,

    and except for the lignin most VS are reactive. The

    digestion process converts part of the VS into gas and

    liquid products, with some solids remaining undigested.

    In the liquid products, VS consist of carboxylic acids,

    extracellular proteins and energy-storage polysaccha-

    rides (Ross, 1998). The following terms are used

    throughout this paper.

    conversion x VS digested

    VS fed9

    yield y total carboxylic acids produced

    VS fed10

    total acid selectivity s

    total carboxylic acids produced

    VS digested11

    total acid productivity p

    total carboxylic acids produced

    L liquid in all reactors time12

    2.13. Statistical analyses

    The statistical analyses were performed using Excel

    software. Analysis of variance for fermentations A and

    D were performed using the student t-test at 5% level

    of significance. The mean and standard deviations of

    the total acid concentrations for fermentations A and

    D were determined from the steady-state operational

    data from days 126 to 257. Correlations between volatile

    solid loading rate (VSLR) and acid productivity (p),

    selectivity (s), yield (y), and conversion (x) were ob-

    tained by fitting experimental results to linear equations

    using Excel software.

    3. Results and discussion

    3.1. Batch reproducibility

    For the following experiments, simulated MSW and

    SS were combined in an 80:20 ratio. To determine the

    reproducibility of MSW/SS batch experiments, two sets

    of three fermentors were operated simultaneously under

    identical conditions. Untreated MSW/SS was used as

    the substrate for one set, whereas lime-treated MSW/

    SS was used for the other. The fermentors contained

    the substrate at a concentration of 100 g dry substrate/

    L liquid. So that the microorganisms would already be

    adapted to the substrate, the inoculum used in these

    experiments was inocula from countercurrent experi-

    ments with the same substrate.

    The initial pH for all fermentations was approximately

    6.5. During the fermentations, the pH varied from 5.4 to

    6.5. The off-gas was analyzed for methane every other

    day, but no methane was found. The combination of

    closure undigested VS dissolved VS carboxylic acids produced biotic CO2 CH4

    mass in water of hydrolysis4

    C. Aiello-Mazzarri et al. / Bioresource Technology 97 (2006) 4756 51

  • 8/8/2019 Aiello Agbogbo y Hotzapple BTech 2006

    6/10

    low pH and iodoform addition effectively inhibited meth-

    anogenesis. Average total carboxylic acids concentra-

    tions for the untreated and lime-treated MSW/SS are

    shown in Fig. 3 along with one-standard-deviation errors

    bars. The variation was less as the beginning and

    increases as the fermentation progresses. The largest

    standard deviation was 0.46 g/L untreated MSW/SS

    and 0.62 g/L for treated MSW/SS. This experiment dem-

    onstrated that batch experiments using lime-treated or

    untreated simulated MSW/SS as substrate system were

    very reproducible.

    3.2. Countercurrent fermentations

    The results from the lime-treated countercurrent fer-

    mentations are shown in Table 2. Fermentations A and

    D were initiated using the same substrate (80% lime-

    treated MSW/20% SS) and the same inocula mixture.

    Both fermentations were started on the same day and

    were operated identically to determine the reproducibil-

    ity of the countercurrent fermentations.

    In Fig. 4, the total carboxylic acid concentration as a

    function of time during fermentations A and D are pre-

    sented. Both fermentations have the same behaviorthrough the operation time (258 d). A t-test was con-

    ducted over the steady-state operation (days 126257),

    and the results showed that there were no significant

    differences between these data set at 5% level of sig-

    nificance. The results show excellent reproducibility for

    countercurrent fermentations.

    Fermentation C had the highest total carboxylic acid

    concentration obtained for lime-treated MSW/SS

    (26.0 g carboxylic acids/L liquid) which surpassed the

    economic goal of 22 g carboxylic acid/L liquid (Holtzap-

    ple et al., 1999). This fermentation was operated at pH

    5.7, LRT = 22.7 d, and VSLR = 8.1 g VS/(L liquid d)

    for 403 d, and also had the highest acid productivity

    (1.36 g total acids/(L liquid d)). The conversion was

    0.302 g VS digested/g VS fed, the yield was 0.18 g car-

    boxylic acid/g VS digested, and the selectivity was

    0.577 g VS digested/g VS fed. Fermentation D (pH

    5.9, LRT = 20.7 d, and VSLR = 3.1 g VS/(L liquid d))

    had the highest conversion (0.449 g VS digested/g VS

    fed). The total carboxylic acid productivity was 1.12 g

    total acids/(L liquid d), the yield was 0.264 g carboxylic

    acid/g VS digested, and the total carboxylic acid produc-

    tivity was 0.577 g total acids/(L liquid d).

    The correlations between volatile solid loading rate

    (VSLR) and acid productivity (p), selectivity (s), yield(y), and conversion (x) are shown in Figs. 58. The re-

    sults show that at higher volatile solids loading rates,

    higher total acid productivities are obtained. Conver-

    sion, selectivity, and yield decrease with increasing

    0

    2

    4

    6

    8

    10

    12

    14

    16

    18

    0 2 4 6 8 10 12 14 16 18 20

    Time (days)

    Totalcarb

    oxylicacid

    concentra

    tion

    (g/L)

    Treated MSW/SS

    Untreated MSW/SS

    Fig. 3. Average carboxylic acid for untreated MSW/SS and treated

    MSW/SS batch fermentations at 100 g substrate/L liquid, error bars 1

    standard deviation.

    Table 2

    Results for lime-treated MSW/SS countercurrent fermentations

    Fermentation trains A B C D E

    Average pH in all fermentors 5.96 0.15 5.79 0.02 5.68 0.06 5.93 0.20 5.85 0.11

    Total carboxyl ic aci d c onc ent ration (g/L) 16. 90 0.77 16.29 1.52 25. 99 1. 46 17.07 0.99 18.62 0. 97

    Acetic acid (wt%) 39.57 4.72 40.55 4.30 40.40 3.10 45.06 4.48 43.89 4.40

    Propionic acid (wt%) 15.70 3.57 14.18 2.10 14.17 1.40 17.81 2.39 12.24 2.21

    Butyric acid (wt%) 23.31 5.00 21.33 3.71 22.15 2.99 18.69 3.63 22.08 3.31

    Valeric acid (wt%) 8.04 1.41 7.77 1.44 6.97 0.96 7.35 1.49 7.25 1.01

    Caproic acid (wt%) 9.47 2.49 12.33 2.96 11.09 2.54 6.98 2.22 9.35 2.82

    Heptanoic acid (wt%) 1.61 0.86 2.64 1.27 1.41 0.74 1.70 1.12 1.45 0.83

    Conversion (g VS digested/g VS fed) 0.432 0.366 0.302 0.449 0.392

    Yield (g total acids/g VS fed) 0.276 0.221 0.175 0.264 0.237

    Selectivity (g total acids/g VS digested) 0.639 0.603 0.577 0.589 0.558

    Total carboxylic acid productivity (g total acids/(L liquid d)) 0.832 0.886 1.360 0.829 1.125

    Biotic CO2 productivity (g CO2/(L liquid d)) 0.390 0.251 0.372 0.290 ND

    Methane productivity (g CH4/(L liquid d)) 0.002 0.001 0.007 0.002 ND

    Mass balance closure (g VS digested/g VS in) 0.96 101.9 100.3 0.93 ND

    Note: All errors are 1 standard deviation.

    ND = not determined.

    52 C. Aiello-Mazzarri et al. / Bioresource Technology 97 (2006) 4756

  • 8/8/2019 Aiello Agbogbo y Hotzapple BTech 2006

    7/10

    VSLR. These correlations are described by the following

    equations:

    p 0:0965VSLR 0:5317 13

    s 0:0090VSLR 0:6274 14

    x 0:0263VSLR 0:5037 15

    y 0:0178VSLR 0:3086 16

    The substrate system, lime-treated MSW/SS, is a hetero-

    geneous lignocellulosic material in which some of the

    components are more digestible by the microorganisms

    (Ross, 1998). MSW is the source of carbon and energy,

    and SS is the source of micronutrients. At high VSLR,

    more substrate is available for the microorganisms,

    which means that more nutrients and easily digested

    substrate are present. The mixed culture of microorgan-

    isms digested the primarily accessible fraction of the

    substrate; however, less carboxylic acids were producedper unit of digested biomass. Part of the digested sub-

    strate is non-acid soluble volatile solids (dissolved VS).

    When substrate and nutrients are abundantly available,

    the microorganisms may produce energy-storage com-

    pounds that can be used in periods of starvation as sug-

    gested by Domke (1999).

    However, at low VSLR, acid productivity was lower,

    but conversion, selectivity, and yield were higher. At

    lower VSLR, less substrate was added, so fewer nutri-

    ents and less easily digestible fractions of the substrate

    were available for the microorganisms. As a conse-

    quence, the microorganisms not only would consume

    the more-available fractions, but also the less-available

    fractions of the substrate. In natural environments, such

    as rumen, substrate concentrations are low, and there-

    fore not enough for the microorganisms to reach maxi-

    mum growth rates (Russell and Baldwin, 1979). At these

    conditions, per unit of digested biomass, the micro-

    organisms may produce more carboxylic acid and less

    energy-storage components.

    The CPDM method developed by Loescher (1996)

    has been used to predict the product acid concentration

    and conversions for countercurrent fermentation sys-

    tems (Ross, 1998; Domke, 1999; Thanakoses, 2002;

    0

    5

    10

    15

    20

    25

    10 30 50 70 90 110 130 150 170 190 210 230 250 270

    Time (days)

    Totalc

    arboxylicacid

    concentration(g/L)

    Fermentation A

    Fermentation D

    Steady-state operation

    Switch to 4 mg iodoform/L

    fed to F4

    8 mg iodoform/Lfed to F4

    Iodoform/Bromoform

    8 mg/L fed to F4

    Fig. 4. Total acid concentration for lime-treated MSW/SS fermenta-

    tion A and D.

    0

    0.5

    1

    1.5

    2

    0 2 4 6 8 10

    VSLR (g VS/(L liquid in all fermentorsd))

    Totalacidproductivity

    (gtotalacids/g/(Ll

    iquidd))

    Experimental data

    Best Line Fit

    Fig. 5. Correlation of total acid productivity with volatile solid

    loading rate.

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    0 2 4 6 8 10

    VSLR (g VS/(L liquid in all fermentorsd))

    Selectivity

    (gtotalacids/gVSdigested)

    Experimental data

    Best line fit

    Fig. 6. Correlation of selectivity with volatile solid loading rate.

    0

    0.1

    0.2

    0.3

    0.4

    0 2 4 6 8 10

    VSLR (g VS/(L liquid in all fermentorsd))

    Yield

    (gtotalacids/gVSfed)

    Experimental data

    Best line fit

    Fig. 7. Correlation of yield with volatile solid loading rate.

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0 2 4 6 8 10

    VSLR (g VS/(L liquid in all fermentorsd))

    Conversion

    (g

    VSdigested/gVSfed)

    Experimental data

    Best line fit

    Fig. 8. Correlation of conversion with volatile solid loading rate.

    C. Aiello-Mazzarri et al. / Bioresource Technology 97 (2006) 4756 53

  • 8/8/2019 Aiello Agbogbo y Hotzapple BTech 2006

    8/10

    Chan, 2002; Aiello-Mazzarri, 2002). CPDM utilizes

    batch data to predict the results of the countercurrent

    fermentations.

    The acetic acid equivalents (Ae) from each of the five

    batch experiment was fit to the equation

    Ae a bt

    1 ct 17

    where t is the time (d) of fermentation, and a, b, and c

    are constants fit by least squares analysis. The rate r

    was obtained from Eq. (18).

    rdAe

    dt

    b

    1 ct2

    18

    The specific rate, r (g Ae produced/(g VS d)) was deter-

    mined from Eq. (18) by dividing it by the initial amount

    of substrate concentration, S0 (g VS/L) in each of the

    five fermentors

    r rS0

    19

    The predicted rate, rpred, was obtained from Eq. (20);

    where the rate of acid production depends on volatile

    solids conversion (x) and product concentration (Ae)

    rpred e1 x

    f

    1 g/Aeh

    20

    Least square analysis was used to determine the empir-

    ical parameter constants e, f, g, and h for rpred (Eq.

    (20)) from the specific rate r(Eq. (19)). Ae was converted

    back to carboxylic acid concentration by / (the ratio of

    total grams of actual acids to grams of Ae). The (1 x)term in the numerator of Eq. (20), is the conversion pen-

    alty function by South and Lynd (1994).

    The conversion, x was calculated using

    xt Ae TLV

    M0 r21

    where, r = selectivity (g Ae produced/g VS digested);

    M0 = the initial substrate mass (g VS).

    The selectivity r, for Eq. (21) was calculated from the

    selectivity s (g total acids produced/g VS digested) deter-

    mined in the countercurrent experiment.

    s /r 22The selectivity (s) used in the CPDM Mathematica pro-

    gram was 0.60 g total acids/g VS digested, the average

    value of selectivity for the lime-treated MSW/SS coun-

    tercurrent fermentations with terrestrial inocula. Selec-

    tivity slightly decreases with increasing VSLR (Fig. 6)

    and was described by Eq. (14). Eq. (20) was used in

    a Mathematica program (Aiello-Mazzarri, 2002) to

    predict acetic acid equivalent concentration (Ae) and

    conversion (x) for the countercurrent fermentation at

    various VSLR and LRT. Ae was converted back to

    carboxylic acid concentration by multiplying by /.

    From the CPDM Mathematica program, the values

    of the predicted total carboxylic acid concentration

    and conversion for the countercurrent fermentation at

    various LRT and VSLR were obtained. CPDM predic-

    tions for conversions and product concentrations for the

    lime-treated MSW/SS countercurrent fermentations

    (substrate = 200 g VS/L liquid) are presented in themap in Fig. 9. The experimental carboxylic acid con-

    centrations and conversions from the lime-treated

    MSW/SS countercurrent fermentations are compared

    with the predicted values from the CPDM map

    shown in Fig. 9. As shown in Table 3, the average abso-

    lute error between the experimental and predicted

    total carboxylic acid concentration and conversion

    for lime-treated fermentations was 4.4% and 15%,

    respectively.

    The error between the predicted conversion, xpred and

    the experimental conversion, xexp as a function of selec-

    tivity is shown in Fig. 10. Linear regression of the data

    gave the following correlation:

    Error xpred xexp

    xexp 2:663s 1:418 23

    Future improvements to the CPDM model should

    incorporate a varying selectivity, r, to better describe

    the countercurrent fermentation process.

    Higher substrate concentrations would be allowed if

    the process were done on a large scale (Holtzapple

    et al., 1999). A higher VS concentration should result in

    higher total carboxylic acid concentrations. CPDM

    was used to predict the conversion and product concen-trations for a VS concentration of 300 g VS/L liquid; the

    CPDM map is shown in Fig. 11. As observed in the

    CPDM map, total acid concentrations as high as

    40 g/L can be reached at LRT of 35 d and VSLR of

    12 g/(L d). Also, conversions as high as 82% can be

    achieved at LRT of 10 d and VSLR of 2 g/(L d). Both,

    high conversions (>70%) and high product concentra-

    tions (>30 g/L) can be achieved at LRT of 35 d and

    VSLR between 2 and 3 g/(L d).

    0

    5

    10

    15

    20

    25

    30

    35

    0 0.2 0.4 0.6 0.8 1

    Conversion (g VS digested/g VS fed)

    Totalcarboxylicacid

    concentration(g/L)

    234

    56812

    10

    15

    2530

    35

    LRT (days)

    VSLR

    (g/(Ld))

    Fig. 9. CPDM map for lime-treated MSW/SS countercurrent

    fermentation (200 g VS/L liquid).

    54 C. Aiello-Mazzarri et al. / Bioresource Technology 97 (2006) 4756

  • 8/8/2019 Aiello Agbogbo y Hotzapple BTech 2006

    9/10

    The CPDM maps for lime-treated MSW/SS coun-

    tercurrent fermentations under mesophilic and thermo-

    philic conditions (Chan, 2002) at 300 g substrate/L

    are compared in Fig. 12. For the same LRT and VSLR

    lower than 8 g VS/(L d), using mesophilic conditions

    (40 C) higher acid concentrations and conversions can

    be achieved than with thermophilic conditions (55 C).

    At high VSLR using thermophilic conditions, acid con-

    centrations as high as 60 g acids/L can be obtained, but

    conversions may be lower than 40%.

    4. Conclusions

    The batch and countercurrent fermentations using

    lime-treated simulated MSW/SS as substrate were very

    reproducible. MSW/SS countercurrent fermentations

    have been well characterized, and countercurrent fer-

    mentations can be operated for extended periods of time

    under steady-state conditions. Addition of 48 mg iodo-

    form/(L liquid added to F4) inhibited methane genera-

    tion. The highest total carboxylic acid concentration

    obtained for lime-treated MSW/SS with terrestrial in-

    ocula was 26.0 g carboxylic acids/L liquid, which sur-

    passed the economic goal of 22 g carboxylic acid/L

    liquid (Holtzapple et al., 1999). The fermentation was

    operated at pH 5.7, LRT = 22.7 d and VSLR = 8.1 g

    VS/(L liquid d) for 403 d. Industrial-scale fermentors

    with higher substrate concentrations would achieve

    higher product acid concentrations and conversions.

    The CPDM model predicts product concentrations

    and conversions within 4% and 16%, respectively. The

    CPDM model was a useful tool in predicting product

    concentration and conversion. The predicted values

    Table 3

    Comparison of experimental and predicted carboxylic acid concentration and substrate conversion for lime-treated MSW/SS countercurrent

    fermentations AF

    Fermentation trains A B C D E F Averagea (%)

    LRT (d) 20.7 19.0 22.7 20.7 17 19.7

    VSLR (g VS/L d) 3.0 4.0 8.1 3.1 6.4 5.7

    Experimental carboxylic acid concentration (g/L) 16.90 16.29 25.99 17.07 18.62 20.66

    Predicted (CPDM) carboxylic acid concentration (g/L) 16.12 17.54 23.53 17.10 18.39 19.97Errorb (%) 4.6 7.7 9.5 0.2 1.2 3.3 4.4Experimental conversion 0.43 0.37 0.30 0.45 0.39 0.33

    Predicted (CPDM) conversion 0.54 0.49 0.29 0.56 0.36 0.38

    Errorb 20.3 24.3 3.3 24.3 7.7 15.1 15.8

    a Average absolute error P

    jpredicted experimental=experimentalj=6.b Error = ((predicted experimental)/experimental) 100.

    0

    5

    10

    15

    20

    25

    30

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    Selectivity

    (g total acid/g VS digested)

    Error

    (%)

    Fig. 10. Correlation of error between experimental and predicted

    conversion and selectivity for lime-treated MSW/SS fermentations

    with terrestrial inocula.

    0

    5

    10

    15

    20

    25

    30

    35

    40

    45

    0 0.2 0.4 0.6 0.8 1

    Conversion (g VS digested/g VS fed)

    Totalcarboxylicacid

    concentration(g/L)

    23

    4568

    12

    10

    15

    2530

    35

    LRT (days)

    VSLR

    (g/(Ld)

    Fig. 11. CPDM map for lime-treated MSW/SS countercurrent

    fermentation for substrate concentration 300 g VS/L liquid.

    0

    10

    20

    30

    40

    50

    60

    0 0.2 0.4 0.6 0.8 1

    Conversion (g VS digested/g VS fed)

    Totalcarboxylic

    acidconcentration(g/L)

    23

    4568

    12

    10

    15

    2530

    35

    346

    Mesophilic

    -------- Thermophilic

    VSLR(g/(Ld))

    LRT (days)

    Fig. 12. CPDM maps for lime-treated MSW/SS four-stage coun-

    tercurrent fermentation under mesophilic and thermophilic conditions

    (80/20 MSW to SS ratio, and 300 g VS/L liquid for both conditions.

    Thermophilic data from Chan, 2002).

    C. Aiello-Mazzarri et al. / Bioresource Technology 97 (2006) 4756 55

  • 8/8/2019 Aiello Agbogbo y Hotzapple BTech 2006

    10/10

    can be used to estimate the costs and process economics

    of industrial-scale fermentors.

    References

    Aiello-Mazzarri, C., 2002. Conversion of municipal solid waste to

    carboxylic acids by anaerobic countercurrent fermentation. Ph.D.

    Dissertation. Texas A&M University, College Station, Texas.

    Bungay, H.R., 1981. Energy, the Biomass Options. John Wiley &

    Sons, New York.

    Caldwell, D.R., Bryant, M.P., 1966. Medium without rumen fluid for

    non-selective enumeration and isolation of rumen bacteria. Appl.

    Microbiol. 14, 794801.

    Chan, W., 2002. Thermophilic anaerobic fermentation of waste

    biomass for producing acetic acid. Ph.D. Dissertation. Texas

    A&M University, College Station, Texas.

    Domke, S.B., 1999. Fermentation of industrial biosludge, paper fines,

    bagasse, and chicken manure to carboxylate salts. Ph.D. Disser-

    tation. Texas A&M University, College Station, Texas.

    EPA, 1998. Municipal solid waste generation, recycling and disposal in

    the United States: facts and figures for 1998. US EPA 530-F-00-024, Washington, DC. Environmental Protection Agency. Avail-

    able from: .

    EPA, 1999. Biosolids generation, use and disposal in the United States,

    1977. US EPA 530-R-99-009, Washington, DC. Environmental

    Protection Agency. Available from: .

    Gizjen, H.J., Zwart, K.B., Teunissen, M.J., Voegels, G.D., 1988.

    Anaerobic digestion of cellulose fraction of domestic residue by

    means of rumen microorganisms. Biotechnol. Bioeng. 32, 749755.

    Hileman, B., 1999. Case grows for climate change. Chem. Eng. News

    77, 1623.

    Holtzapple, M.T., Lundeen, J.E., Sturgiss, R., Lewis, J.E., Dale, B.E.,

    1992. Pretreatment of lignocellulosic municipal solid waste by

    ammonia fiber explosion (AFEX). Appl. Biochem. Biotechnol. 34,

    521.

    Holtzapple, M.T., Ross, M.K., Chang, N.S., Chang, V.S., Aldelson,

    S.K., Brazel, C., 1997. Biomass Conversion to Mixed Alcohol

    Fuels Using the MixAlco Process. In: Saha, B.C., Woodward, J.

    (Eds.), ACS symposium series 666. ACS, Washington, DC, pp.

    130142.

    Holtzapple, M.T., Davison, R.R., Ross, K., Aldrett-Lee, S., Nagwani,

    M., Lee, C.M., Lee, C., Adelson, S., Karr, W., Gaskin, D., Shiraga,

    H., Chang, N.S., Chang, S., Loescher, M., 1999. Biomass conver-

    sion to mixed alcohol fuels using the MixAlco process. Appl.

    Biochem. Biotechnol. 7779, 609631.

    Klass, D.L., 1998. Biomass for Renewable Energy, Fuels, and

    Chemicals. California Academic Press, San Diego, pp. 651653.

    Loescher, M.E., 1996. Volatile fatty acids fermentation of biomass and

    kinetic modeling using the CPDM method. Ph.D. Dissertation.

    Texas A&M University, College Station, Texas.

    Ross, M.K., 1998. Production of acetic acid from waste biomass.

    Ph.D. Dissertation. Texas A&M University, College Station,

    Texas.

    Russell, J.B., Baldwin, R.L., 1979. Comparison of substrate affinities

    among several rumen bacteria: a possible determinant of rumen

    bacterial competition. Appl. Environ. Microbiol. 37, 531536.

    South, C.R., Lynd, L.R., 1994. Analysis of conversion of particulate

    biomass to ethanol in continuous solids retaining and cascade

    bioreactors. Appl. Biochem. Biotechnol. 4546, 467481.

    Sterzinger, G., 1995. Making biomass energy a contender. Technol.

    Rev. 98, 3440.

    Thanakoses, P., 2002. Conversion of bagasse and corn stover to mixed

    carboxylic acids using a mixed culture of mesophilic microorgan-

    isms. Ph.D. Dissertation. Texas A&M University, College Station,

    Texas.

    56 C. Aiello-Mazzarri et al. / Bioresource Technology 97 (2006) 4756

    http://www.epa.gov/garbage/pubs/mswfinal.pdfhttp://www.epa.gov/oigearth/ereading_room/BIOSOLIDS_FINAL_REPORT.pdfhttp://www.epa.gov/oigearth/ereading_room/BIOSOLIDS_FINAL_REPORT.pdfhttp://www.epa.gov/oigearth/ereading_room/BIOSOLIDS_FINAL_REPORT.pdfhttp://www.epa.gov/oigearth/ereading_room/BIOSOLIDS_FINAL_REPORT.pdfhttp://www.epa.gov/garbage/pubs/mswfinal.pdf