You are on page 1of 8

Int. J. HydrogenEnergy, Vol. 17, No. 4, pp. 281-288, 1992. 0360-3199/92 $5.00 + 0.

00
Printed in Great Britain. Pergamon Press Ltd.
International Associationfor Hydrogen Energy.

BIOLOGICAL CONVERSION OF SYNTHESIS GAS INTO FUELS

K. T. KLASSON,C. M. D. ACKERSON, E. C. CLAUSEN and J. L. GADDY


Department Chemical Engineering, University of Arkansas, Fayetteville, AR 72701, U.S.A.

(Received for publication 3 January 1992)

Abstract--Liquid and gaseous fuels may be produced from coal biologically by the indirect conversion of coal synthesis
gas. Methane has been produced from synthesis gas using acetate and CO2/H2 as intermediates, utilising a number of
CO-utilisingl and methanogenic bacteria. Also, a bacteria has been isolated from natural inocula that is capable of produc-
ing ethanol from synthesis gas through indirect liquefaction. This presentation summarises the research to optimise the
performance of these cultures. These conversions, involving H2 and CO which are only slightly soluble, are severely
mass transfer limited, and methods to enhance mass transport are examined. Experimental results and models for several
reactor designs, including CSTR and packed columns, are presented and discussed.

INTRODUCTION The purpose of this paper is to present data for the


development of optimal bioreactor concepts for syngas
Synthesis gas, a mixture of primarily CO, H2 and CO2, is
fermentations. Laboratory data for continuous culture
a major building block in the production of fuels and
experiments for the conversion of synthesis gas components
chemicals. The gas may be produced from several sources,
into methane and ethanol are presented. Various bioreactor
including coal, oil shale, tar sands, heavy residues, biomass
schemes for synthesis gas fermentations have been
or natural gas. Most synthesis gas is produced today by
investigated and mathematical models that define intrinsic
catalytic reforming of natural gas, although the partial
kinetics and mass transfer relationships are developed.
oxidation of heavy liquids is also practiced [1]. Only a
Methods to predict reactor performance and gas retention
small percentage of the synthesis gas currently produced is
times for the CSTR and immobilised cell reactor are
by gasification of solid fuel. However, because of the large
presented.
reserves of coal in the U.S.A. (300 year supply at the cur-
rent consumption rate [2]), synthesis gas production from
coal will become an important technology in the future. SYNTHESIS GAS FERMENTATIONS
Coal gasification, which is a combination of pyrolysis
and combustion reactions [ 3 ], produces a gas consisting of
more than 50% H2 and CO, the balance being a mixture of Methane production
CO2, CH4, H2S, COS and nitrogen compounds. The actual Methane may be produced by methanogenic bacteria
composition depends upon process conditions and the coal from either acetate or H2 and CO2, both of which may be
that is employed. The raw gas has a low to medium Btu produced from syngas components. Acetate may be pro-
content, with a heating value of 1 6 0 - 4 5 0 Btu SCF t, duced by several anaerobic bacteria, including Peptostrep-
depending on whether air or oxygen is used during gasifica- tococcus productus [6, 7 ], Acetobacterium woodii [ 8],
tion [4]. Following quenching and purification, the syn- Clostridium thermoaceticum [9, 10,] and Eubacterium
thesis gas contains 2 5 - 3 5 % H2, 4 0 - 6 5 % CO, 1 - 2 0 % limosum [ 11 ], which produce acetate by the reaction:
CO2, 0 - 7 % CH4 and other compounds in small quan-
tities. 4 CO + 2 H20 ~- CHsCOOH + 2 CO2. (1)
Catalytic processes may be used to convert syngas into a
variety of fuels and chemicals, such as methane, methanol, Among these bacterial species, P. productus utilises CO
formaldehyde, acetic acid, etc. [5]. Microorganisms may very rapidly with a doubling time of less than 2 h, and can
also be used to convert synthesis gas components into fuels grow with as much as 90% CO in the gas phase [6].
and chemicals. Biological processes, although generally Many anaerobic bacteria, including P. productus, are
slower than chemical reactions, have several advantages known to produce acetate from H2 and CO2 [ 1 2 - 1 4 ] ,
over catalytic processes, such as higher specificity, higher which produces a homoacetic fermentation by anaero-
yields, lower energy costs and generally greater resistance bically oxidising hydrogen and reducing CO., according to
to poisoning. Furthermore, the irreversible character of the equation:
biological reactions allows complete conversion and avoids
thermodynamic equilibrium relationships. 4 H2 + 2 CO2 ~- CH3COOH + 2 H20. (2)

281
282 K.T. KLASSON et al.

Two species of purple non-sulfur bacteria, Rhodo- tion by R. rubrum is essentially unaffected by CO partial
pseudomonas gelatinosa [ 15, 16] and Rhodospirillum pressures up to 2.0 atm. Therefore, the limiting factor in
rubrum [17], are known to perform the water gas shift CO utilisation by R. rubrum is the ability to maintain a high
reaction to produce H2 as follows: cell concentration and, consequently, a low dissolved CO
tension in the liquid phase. M. formicicum and M. barkeri,
CO + H20 ~ H2 + CO2. (3) on the other hand, have been shown to be susceptible to CO
inhibition at partial pressures of only 0.76 atm. Conse-
R. gelatinosa grows under strict anaerobic conditions in the quently, a high cell concentration of R. rubrum in the co-
dark with CO as the only carbon and energy source, culture will be essential to avoid inhibition of the
although growth is stimulated by the addition of trypticase. methanogens.
R. rubrum requires tungsten light and the presence of a R. rubrum is a photosynthetic bacteria requiring tungsten
carbon source other than CO (sugars, acetate, yeast extract light for growth, but not for CO uptake. Figure 1 shows the
etc.) for growth. In comparing these two species, R. growth and consumption of CO with time at various light
rubrum grows faster and reaches higher cell concentrations intensities for R. rubrum. As shown, the cell growth rate
that uptake CO more rapidly. R. rubrum has also been increases with light intensity up to 1490 iux, however, no
found to tolerate small amounts of oxygen and sulfur com- further enhancement is found at higher intensities. CO con-
pounds often present in synthesis gas. sumption was essentially unaffected by the presence of
Almost all methanogenic bacteria, including Methano- light. Methanogens have been found to be unaffected by the
spirillum hungatii, Methanobacterium formicicum, presence of light.
Methanobrevibacter smithii, Methanosarcina barkeri, An experiment was performed in a continuous stirred
u t i l i z e CO2 and H2 to produce CH4 according to tank reactor to study the simultaneous conversion of CO2
[ 18-20] : and H2 directly to CH4 employing a co-culture of R.

4 H2 + CO2 ---~-CH4 + 2H20. (4)


500
I Light
Methane may also be produced from acetate by Methano- ~ / ~ rntens,ty __
sarcinaceae, such as Methanosarcina barkeri, as well as --} 400 / \% / • 2720 Jux
Methanothrix soehngenii [21 ] according to: • o 1490 lux
CH3COOH = CH4 + CO2. (5) }300 ~ /. /...4~ • 6701ux
// / "\,• i ~ 0 lux
While M. barkeri will utilise acetate only in the absence of - 200
other preferred substrates (such as H2 and COJ,
Methanothrix sp. does not utilise H2 and CO2 and growth
and methane formation is observed exclusively in the
presence of acetate [22 ]. Both microorganisms show com-
parable specific growth rates at low acetate concentrations
( < 3 raM). However, from the Monod saturation constants 0 20 40 6C ~C, 100 ~2C, 140 160

(Ks = 0.7 mmol 1-~ for Methanothrix and 5 mmol 1-~ for Time (hrs)
M. barked), it is expected that at low acetate concentrations
Methanothrix will give faster rates and predominate. Fig. I. Dry cell weight concentration ofR. rubrum at vaarious light
intensities.
From the above, it can be seen that the production of
methane from syngas is a two- step process; formation of
the methane precursors (acetate or hydrogen) and the 2.0
biomethanation of the precursor. These reactions may be
carried out in separate stages or as a co-culture in the same
1.5 /
reactor. Compatibility of the cultures with substrates and
products is essential for an efficient process. It has been
found that the methanogenic bacteria are highly sensitive to
1.O
low concentrations of acetic acid ( > 1 2 g I-~). This
inhibition results in low productivities and a large second
stage reactor. Consequently, the acetate pathway is not con- /
0.5 /
sidered as promising as the hydrogen pathway.

Methane production from H2 and CO:. Methane may be 0.0 8 l __ l


0 I00 200 300 400 500
produced from synthesis gas by conversion of CO and Hz
by R. rubrum, followed by conversion of all the H2 and an Time (hrs)
equivalent amount of CO2 to methane using either M. for- Fig. 2. Rate of methane production by a co-culture of R. rubrum
micicum or M. barkeri. It has been found that H2 produc- and M. formicicum in the CSTR.
BIOLOGICAL CONVERSION OF SYNTHESIS GAS 283

EztO00So0;
+tC0t+(%o,//@
005; O.)
rubrum and M. formicicum. Since the organisms have dif- 1.5
ferent optimum temperatures, the lower temperature, Yellt

30°C, was chosen for study. Figure 2 shows the CH4 and
CO production with time since since start up in the CSTR.
Following a significant period of methanogen acclimation,
almost complete conversion of both CO and H2 occurred
after 300 h of operation. The methane production rate "~
o 10
shown in Fig 2 reached a steady state level after 350 h of
operation of about 1.6 mole CH4 h- t, which represents a E
:o+ /.
methane yield from CO and H: and COz of about 96% of v
theoretical. The system was operated with a retention time ,,
of 1 h and stable operation was monitored for several
weeks. 2: O.S

Ethanol production
While many anaerobic, facultatively anaerobic and even
some strictly aerobic microorganisms form various
amounts of ethanol from glucose [23], no organisms were
"t®
to
known to form ethanol autotrophically from synthesis gas 0.0 . . . . i . . . . i . . . . . .

components. In 1987, a strict anaerobic mesophilic 0 5 I0 15


bacterium was isolated that was capable of converting CO,
H~ and CO2 to a mixture of acetate and ethanol [24n]. N^c . (mmol)
Identification and characterization studies have shown Fig. 3. Product distribution for C. Ijungdahlii with various
that the bacterium is a new clostridial species, named initial yeast extraction concentrations.
Clostridium ljungdahlii, Strain PETC, in honor of Dr Lars
G. Ljungdahl for his work on clostridia and acetogens molar ethanol to acetate ratio of 0.10 was obtained for yeast
[25 ]. C. lungdahlii is a gram-positive, motile, rod-shaped extract concentrations between 0.005 and 0.05%. The
anaerobic bacterium which sporulates infrequently. In addi- normal ratio of 0.05 results under more favourable growth
tion to synthesis gas components, it is capable of growing conditions when employing 0.1 and 0.2% yeast extract.
on xylose, arabinose and fructose. As with other class I Studies with a defined medium of only vitamins, minerals
clostridia, it is expected that ethanol and acetate are formed and salts showed similar results in increasing the product
through acetyl-CoA as the central intermediate [26]. ratio.
The overall stoichiometry for the formation of ethanol Recent research has shown that the presence of reducing
and acetate from CO and HdCO2 has been established as agents in the liquid media of Clostridium fermentations has
[27]: brought about an increase in solvent formation [28,29].
Reducing agents apparently cause altered electron flow,
6 CO + 3 H20 -- CH3CH2OH + 4 CO2 (6) which directs carbon flow and acid to alcohol production.
2 CO2 + 6 H2 ,. CH3CH2OH + 3 H20 (7) Reducing equivalents are directed to the formation of
NADH which, in turn, result in increased alcohol produc-
4 CO + 2 H20 ,- CH3COOH + 2 CO_+ (8) tion. Batch experiments were carried out with C. ljundahlii
2 CO2 + 4 H2 --- CH3COOH + 2 H20. (9) by adding small quantities of reducing agents (30, 50 and
100 ppm) to assess the feasibility of increasing the ethanol
Under usual laboratory conditions, C. Ijungdahlii pro- to acetate ratio. The experiment with 100 ppm of reducing
duces acetate as the major product, with an ethanol/acetate agents resulted in very limited growth in all cases. On the
ratio of only 0.05. It was noted that yeast extract has an other hand, 50 and 30 ppm concentrations were successful
influence on the product ratio, which led to the hypothesis in improving the ethanol to acetate ratio in some cases, as
that higher ethanol production is non growth related. An is shown in Table 1. The experiment with benzyl-viologen
examination of the acetyl CoA pathway shows that produc- at a concentration of 30 ppm produced 3.7 mmol of ethanol
tion of acetate is balanced in ATP, while ethanol production with a ratio of 1.1, the highest ratio observed in batch
results in a net consumption of ATP, which would not sup- experiments. It is interesting to mention that those reducing
port growth of the bacteria. Therefore, studies to minimise agents that improved the product ratio always resulted in
acetate production have concentrated upon factors which slower growth rates of the bacteria, as could be expected
regulate the growth of the organism. from decreased ATP formation.
Recently, the connection between sporulation and
Control of growth rate parameters. C. ljungdahlii grows increased solventoganesis has been identified [ 3 0 - 3 2 ] .
well and produces ethanol and acetate within a pH range of Under certain conditions, which are strain-dependent, a
4 - 6 with typical anaerobic media. Figure 3 shows the pro- shift of the bacteria into a sporulation phase is accompanied
duct distribution for C. ljungdahlii with various initial by morphological changes (elongation of the cells) and the
yeast extract concentrations in batch culture. As noted, a production of solvents rather than acids. A batch experi-
284 K.T. KLASSON et al.

Table I. Peak levels for ethanol production and the molar ratio (ETOH/ACH) at
30 and 50 ppm reducing agent concentrations

50 ppm 30 ppm

Reducing agent ETOH ETOH/ACH ETOH ETOH/ACH


(mmol) (mmol)

Control 0.60 0.12 1.40 0.24


Sodium thioglycolate 1.30 0.20 1.30 0.25
Ascorbic acid 1.50 0.24 1.50 0.25
Menthyl viologen 1.90 0.20 2.50 0.40
Benzyl viologen 1.25 0.21 3.70 1.10

ment with C. ljungdahlii was conducted on the premise that time, controlling the growth rate. Media constituents to
by forcing the culture to grow at a reduced rate, sporulation promote growth can be added to the first reactor, and con-
could be induced with an accompanying improvement in stituents to promote ethanol production at the expense of
ethanol production. Synthesis gas was used as the primary acetate can be added to the second reactor.
carbon substrate. However, the complex nutrient, yeast Figure 4 shows the molar product ratios for both stirred-
extract, was replaced by various sugars and starches which, tank reactors. Yeast extract (0.02 %) was added to the liquid
in previous studies, promoted sporulation of Clostridium medium of Reactor A (first in the series) initially and
thermosaccharolyticum [ 33 ]. cellobiose later. Ethanol concentrations in Reactor B
Table 2 summarises the results obtained for each of the
nutrients studied, along with the maximum values obtained
for cell concentration, ethanol concentration and molar pro- 1.5
YEAST E X T R A C T CELLOBIOSE
duct ratios. As noted, the highest product ratios were
obtained for cellobiose and rhamnose, with product ratios 1- /
over 3 times the ratio obtained in the presence of yeast 0 -"q 4,~'
I-
extract. Ethanol and cell concentrations were highest in the LU . , i//
presence of cellobiose and galactose, where the ethanol
concentrations were over 4 times the value obtained in the

1.0 ,'¢"-- e.. ~ . .e"
/,
'~5--~ ~

presence of yeast extract and the cell concentrations were


20% greater. Thus, cellobiose as a nutrient produces not '5
only higher ratios of ethanol to acetate, but also higher
concentrations of ethanol and cells. 2 , i ~ Switch to I

Continuous stirred-tank reactor performance. An


F~ACTOA B
J
obvious method to produce high ethanol ratios is to operate t L

0 2 4 6 8 10 12 14
two continuous reactors in series, with the first used to pro-
mote cell growth, while the second reactor is used for Time (days)
increased ethanol production. A pH shift between the reac-
tors from 4.5 to 4.0, as well as a dilution rate shift, are used Fig. 4. Molar product ratios attained in a two-stage CSTR system
to cause the onset of ethanol production while, at the same with C. ljungdahlii

Table 2. Summary of results with nutrient sources bringing about sporulation

Maximum

Cell conc. ETOH ETOH/ACH


Nutrient (mg 1 ~) (mmol) molar ratio

Yeast extract 140 0.13 0.13


Cellobiose 170 0.56 0.45
Rhamnose 135 0.31 0.44
Galactose 168 0.53 0.36
Starch 130 0.27 0.36
BIOLOGICAL CONVERSION OF SYNTHESIS GAS 285

increased to nearly 3 g 1-~ and seemed to be stimulated resistance around the cells is usually neglected with respect
somewhat by the use of cellobiose as the nutrient for cell to other resistances, because of the minute size and tile
growth. Substrate CO and H2 conversions were essentially enormous total surface of the cells [37]. Thus, for the
100% in Reactor A, and fluctuated somewhat in Reactor B. transfer of sparingly soluble gases, such as CO, the primary
The product ratio increased with time in both reactors, resistance to transport may be assumed to be in the liquid
reaching a value of about 1.0 in Reactor A and a value of film at the gas-liquid interface.
about 1.5 in Reactor B. The addition of cellobiose seemed It can be shown that the substrate transfer rate per unit
to improve the product ratio over yeast extract. By subtrac- of reactor volume, d~sX/VLdt is given in terms of the gas
ting the product concentrations produced in Reactor A, an phase partial pressures as:
ethanol ratio of 4 mol/mol is obtained in Reactor B. The
specific productivity steadily improved to levels of dN~ KLa
2 5 0 - 3 0 0 mole ethanol g cell ~ day -t throughout the (Psc - PsL) (lO)
VLdt H
experiment, which is a 30-fold improvement over specific
productivities in a single CSTR.
where Ns~ = moles substrate transferred from the gas
phase, is the volume of the liquid phase, t is time, KL is
BIOREACTOR DESIGN the overall mass transfer coefficient, a is the gas-liquid
interfacial area per unit volume, H is Henry's law constant,
The choice of a suitable bioreactor for synthesis gas
PsG is the partial pressure of the substrate in the the bulk
fermentations will be a matter of matching reaction kinetics
gas phase, and PsL is the partial pressure of (dissolved
with the capabilities of the various reactors. It has been
tension) of the substrate in the liquid phase (Ps~ = HCL).
found that for these slightly soluble gases, the rate of mass
The rate of transport from the gas phase must be equal to
transfer usually controls the reactor size [34,35]. Mass
the rate of consumption in the liquid phase, given by a
transfer capabilities of the reactor must be balanced with
Monod relationship:
the cell density achieved. The proper reactors for these
systems will likely be ones that achieve high mass transfer"
d~s Xqm ~ L KL a
rates and high cell densities. These concepts will be (PsG - P ~ ) (ll)
expanded in this section. VLdt L 2
K~ + PsL + (Ps) /W i
H

Gas-liquid mass transfer concepts where X is cell concentration and qm, g~ and W' are
The transfer of gas phase substrates in fermentation systems Monod constants.
involves three heterogeneous phases: the bulk gas phase, Equation (11) shows that a bioreactor for these gaseous
the culture medium (liquid) and microbial cells (solid) systems must operate in either of two regimes. In one case,
suspended in the medium. The reactants, present in the gas sufficient cells are present to react more solute, but the
phase, must be transported across the gas-liquid interface mass-transfer rate cannot keep pace. Therefore, the liquid
and diffuse through the culture medium to the cell surface phase concentration goes to zero and the reactor is mass
to be consumed by the microbes. In general, a combination transport limited. The cell concentration and rate of con-
of the following resistances can be expected [36]: sumption are limited by the ability of that particular reactor
to transfer substrate. In the other case, sufficient substrate
(1) diffusion through the bulk gas to the gas-liquid can be supplied, but the cell concentration does not allow
interface; consumption at an equal rate. Then the liquid phase concen-
(2) movement across the gas-liquid interface; tration is not zero (with possible inhibitory effects) and the
(3) diffusion of the solute through the relatively unmixed rate is limited by the cell concentrations in that particular
liquid region (film) adjacent to the bubble and into bioreactor. Obviously, the best bioreactor is one that will
the well-mixed bulk liquid; achieve high cell concentrations and high mass transfer
(4) transport of the solute through the bulk liquid to the rates.
stagnant film surrounding the microbial species;
(5) transport through the second unmixed liquid film Bioreactors for synthesis gas fermentations
associated with the microbes; Since large volumes of syngas must be processed, con-
(6) diffusive transport across the liquid/solid boundary tinuous reactors are dictated. Stirred-tank reactors achieve
and into the microbial floc, mycelia, or particle, if high mass transfer rates, but require substantial energy
appropriate. When the microbes take the form of input for agitation. Immobilised cell reactors achieve high
individual cells, this resistance disappears; cell concentrations, without agitation, and are promising
(7) transport across the cell envelope to the intracellular for these applications. Trickle-bed columns, where the gas
reaction site. is the continuous phase and the liquid flows over packed
internals, is a unique means of increasing the mass transfer
As with the conventional chemical engineering analysis for these systems.
of absorption processes, mass transfer through the bulk gas
phase is assumed to be instantaneous. Also, when Stirred tank reactor. The traditional CSTR assumes com-
individual cells are suspended in a medium, the liquid film plete mixing and uniform concentrations throughout the
286 K.T. KLASSON et al.

bulk liquid phase. For syngas fermentations, the gas must with a linear relationship. The slope of this line gives the
be sparged into the liquid phase, be consumed, with any mass transfer coefficient, KLa/H = 30. A model including
excess and product gases leaving the top of the liquid and equation (12), as well as material balances for the gases
eventually the reactor. High gas flow rates are required and flowing into the reactor and equilibrium relationships for
near complete conversion of substrate is necessary. Con- the gas phase CO2 with the bicarbonate and the pH level in
versely, only small liquid flow rates, essential to supply the liquid, has been developed [ 35]. Solutions of the model
nutrients and remove liquid products, are necessary. Con- for various volumetric mass transfer coefficients and
sequently, high cell concentrations should be possible. In various total operating pressures are shown in Figs 6 and
most cases, the reactor volume will be controlled by the 7, respectively. Experimental data at 1 atm and a mass
necessary gas retention time to achieve the desired conver- transfer coefficient of 30 are also included in the figures.
sion of substrate. Relatively high agitation rates will be As observed, increases in the mass transfer coefficient or
required to promote transfer of the slightly soluble gas in total operating pressure lead to higher reactor produc-
substrate. tivities. Due to the perfect mixing in a CSTR, complete
Mass transfer coefficients, necessary for prediction of conversion is only possible when the gas flow rate is very
CSTR performance and scale-up, may be obtained from an low.
analysis of the operation under mass-transfer limited condi- Figure 6 shows that with a mass transfer coefficient of
tions. A material balance around the CSTR with perfect 100, a pseudo retention time of 1 h would result in a con-
mixing gives the relationship defining concentrations: version of 80%. From Fig. 7, the retention time could be
reduced to 6 min at 10 atm for the same amount of CO con-
verted. The use of the model allows the extrapolation of
1 1 VL KLa P~
- + (12)
Yo E E H nl !00 j K=a/H = 100
L

Equation (12) is expressed in terms of an inert component, 80 i //'/S--/


whose quantity and partial pressure does not change
through the system. Therefore, to the model, concentra- ~ 60 ,'- /'
tions are in the ratio of substrate to inert (Y), with Y,, at
~ L a / H = 10
the outlet and E at the inlet. P~ is the partial pressure of 40 ///' / ~
inert, n the gas stream and n~ is the molar flow rate of o
inert. 0
The agreement of this model with experimental data for P. o
productus is shown in Fig. 5. Good agreement is achieved

0 1 2 3 4
1.5 Pseudo Retention Time (hrs)

Fig. 6. Syngas conversion as a function of pseudo retention time


and mass transfer coefficient in a CSTR.

10 140
8
>. "U
e.- -2o~ t
100 X , " 10

/•
-J

O
¢.)
05 • ExpE~lrr,Lwltal 80 "'\\
Data \,\
o Inlet C*,a5
E

S
O
20 P, = 1 |
0 . 0 . . . . . . . . • . . . . . t _ _ ~ _ _ . ~ . . . .

o.0o 0.0 1 0.02 0.03 0.04


05 10 ~5 20 25 30 35 4.0 45 5.0
!
V L P~ / Y co nl
Pseudo Retention Time (hrs)
Fig. 5. Model verification and evaluation of the volumetric mass Fig. 7. Productivity as a function of total pressure and pseudo
transfer coefficient in a CSTR. retention time in a CSTR.
BIOLOGICAL CONVERSION OF SYNTHESIS GAS 287

performance of the CSTR system and will permit 12o I


preliminary economic evaluation of an industrial scale i KkIII:JH -- 1 0 0
process when coupled with suitable equations for scale up
of properties such as the mass transfer coefficient.
e" 6 C I- ' / J J
o
lmmobilised cell column (ICR). Column fermenters, with "~ ; , °° ° et/H " 13
immobilised or suspended cells, offer the advantages of
high cell densities and plug flow operation. These systems 0 i ' , el"
do not require mechanical agitation, with mixing provided 0 -~c '- I'i" ¶/
by counter flow of gas and liquid. Energy for mixing is sup- 0 ~ /' / /
tO 2,S I /
plied by gas pressure drop and such systems are potentially ,

more economical than the CSTR. Packed columns also


© r i i i
offer the advantages of high surface to volume ratios and
O 1 2 3 4
high mass transfer rates with reduced back mixing.
Whole cell immobilisation techniques can be classified Pseudo Retention Time (hrs)
into two major categories; entrapment and carrier binding
Fig. 8. Syngas conversion as a function of pseudo retention time
[38]. Entrapment includes both enclosure of a catalyst and mass transfer coefficient in an ICR.
behind a membrane or within a gel structure. Carrier bin-
ding includes all methods where there is a direct binding of
cells to water insoluble carriers by physical adsorption or umn, R is the ideal gas constant, T is the absolute tem-
by ionic and/or covalent bonds. Potential mass transfer perature, and G is the gas flow rate. A plot of In E/T,>
limitations are always present with entrapment systems, vs ShRT/G yields a straight line with slope KLaes/H. The
either across the gel matrix or gel occlusion, or across the numerical solution (Runge-Kutta) of the differential equa-
system membrane. On the other hand, the carrier binding tions that describe the system were solved for other
methods allow direct contact between the fermentation operating conditions and are shown in Fig. 8. Experimental
broth and the biocatalyst, with potentially enhanced mass data are given for KtaEL/H of 13.5.
transfer rates. The immobilised cell column achieves higher rates of
Microorganisms can be immobilised to insoluble biosup- specific CO conversion than the stirred tank reactor without
port materials by two methods: cross linking and adsorp- the need for more expensive mechanical agitation. More
tion. Crosslinking, or covalent bonding, involves the use of importantly, at the same mass transfer coefficients as in the
a chemical agent, like glutaraldehyde or cyanuric chloride, CSTR, conversions are substantially higher. For example,
to link the cells to the support. The chemical reaction is bet- at KLa/H= 100, the conversion at a 1 h retention
ween the hydroxyl or lipid groups in the cell wall and a time is 95%, compared to 80% for the CSTR. Alter-
durable coating, like gelatin or agar, applied to the packing. natively, 80% conversion could be achieved in a retention
Adsorption is the physical (occasionally ionic) attachment time only 3 min in the ICR. The major disadvantage of the
of the cell to the support. This method has been found to ICR is the lack of flexibility in operating conditions since
be effective for some small bacteria that can adhere to the contacting capabilities are fixed by the column and
crevices in a support like wood chips. packing dimensions.
In these reactors, the microorganisms are in direct
contact with the substrate, minimising diffusional CONCLUSIONS
resistance. These packed columns operate close to plug The efficient fermentation of coal synthesis gas to
flow and, thereby, offer kinetic advantages for these reac- methane and ethanol has been demonstrated. Two pathways
tions. Cells attached to the support grow and multiply into for the indirect production of methane from synthesis gas
a film, which may be several layers of cells in thickness. have been evaluated. Production through acetate as an
In fact, cell overgrowth can result in completely filling the intermediate is limited by acetate inhibition of methano-
interstitial spaces, such that channeling may be a problem. gens. Production through H2 with a co-culture of R.
Therefore, high cell densities and low retention times are rubrum and methanogens gives faster rates without inhibi-
possible. tion.
By combining a material balance along the column with Ethanol can be produced from synthesis gas with a new
the rate expression for gas transport into the liquid phase, species of Clostridium isolated from animal waste. The
the following expression for the ratio of partial pressures of ratio of ethanol to acetate in the product stream is affected
gaseous reactant entering and leaving the reactor is by many variables including pH, nutrient composition and
obtained: the introduction of reducing agents to alter electron flow.
High ethanol ratios are favoured by non-growth conditions.
P~' KLa ELh RTS
In pi - (13) Product ratios of 4:1 (ethanol to acetate) are achieved in a
s H G two-stage continuous culture with pH and dilution rate
shift.
where eL is the fraction of liquid in the column, h is the Bioreactors that achieve high mass transfer rates and high
height of the column, S is the cross sectional area of the col- cell concentrations are desirable for synthesis gas fermenta-
288 K, T. KLASSON et al.

tions. Methods to determine mass transfer coefficients for 18. R. K. Thauer, K. K. Jungnermann and K. Decker, Energy
CSTR and ICR reactors have been developed. High conservation in chemotrophic anaerobic bacteria. Bacteriol.
pressure has been found to increase the reaction rate pro- Rev. 41, 100-180 (1977).
portionately. Models for these bioreactors show high con- 19. W. E. Balch, G. E. Fox, L, J. Margum, C. R. Woese and R.
S. Wolfe, Methanogens: reevaluation of a unique biological
version o f gaseous substrate can be achieved in a retention
group. Microbiol. Rev. 43, 260-296 (1979).
time o f a few minutes. 20. A. J. B. Zehnder, K. Ingvorsen and T. Marti, Microbiology
of methane bacteria, in Hughes et al. (eds), Anaerobic Diges-
REFERENCES tion. Elsevier Biomedical Press, Amsterdam
21. W. J. Jones, D. P. Nagle, Jr and W. B. Whitman,
1. M. S. Graboski, The production of synthesis gas from Methanogens and the diversity of archaebacteria. Microbiol.
methane, coal and biomass. In R. G. Herman (ed.), Catalytic Rev. 51, 135-177 (1978).
Conversion of Synthesis Gas and Alcohols to Chemicals, 22. B. A. Huser, K. Wuhrmann, and A. J. B. Zehnder,
pp. 37-50. Plenum Press, New York (1984). Methanothrix soehgenii gen. nov. sp. nov., a new
2. R. Specks and A. Klussmann, German hard coal conversion acetotrophic non-hydrogen-oxidizing methane bacterium.
projects. Energy Progr. 2, 6 0 - 6 5 (1982). Arch. Microbiol. 132, 1 - 9 (1982).
3. D. R Simbeck, R. L. Dickenson, A. J. Moll, Coal gasifica- 23. J. Wiegel, Formation of ethanol by bacteria. A pledge for the
tion, an overview. Energy Progr. 2, 4 2 - 4 6 (1982). use of extreme thermophilic anaerobic bacteria in industrial
4. J. M. Coffin, Industrial coal gasification: applications and ethanol fermentation processes. Experientia 36, 1434-1446
economy. Energy Progr. 4, 131 - 137 (1984). (1980).
5. Ph. Courty and P. Chaummette, Syngas: a promising 24. S. Barik, S. Prieto, S. B. Harrison, E. C. Clausen and J. L.
feedstock in the near future. Energy Progr. 7, 23 - 3 0 (1987). Gaddy, Biological production of ethanol from coal synthesis
6. S. Berik, E. R. Johnson, E. C. Clausen and J. L. Gaddy, Con- gas. Biotechnology Applied to Fossil Fuels. CRC Press,
version of coal synthesis gas to methane. Energy Progr. 7, Boca Raton (1987b).
157 (1987a). 25. E. C. Clausen and J. L. Gaddy, Advanced studies of the
7. W. H. Lorowitz and M. P. Bryant, Peptostreptococcus pro- biological conversion of synthesis gas to methane. Topical
ductus strain that grows rapidly with CO as the energy source. Report 1: Reactor Optimization, Performed on METC Con-
Appl. Environ. Microbiol. 47, 961-964 (1984). tract DE-AC21-86MC23281, U.S. Department of Energy
8. R. Kerby, W. Niemczura and J. G. Ziekus, Single carbon (March 1988).
catabolism in Acetobacterium woodii and Butyribacterium 26. P. Rogers, Genetics and biochemistry of clostridium relevant
methylotrophicum by fermentation and ~3C nuclear magnetic to development of fermentation processes. Adv. Appl
resonance measurement. J. Bacteriol. 115, 1208-1218 Microbiol. 31, 1 - 6 0 (1986).
(1983). 27. J. L. Vega, S. Prieto, B. B. Elmore, E. C. Clausen and J. L.
9. H. G. Wood, H. L. Drake and S. Hu, Studies with Gaddy, The biological production of ethanol from synthesis
Clostridium thermoaceticum and the resolution of the pathway gas. Appl. Biochem Biotechnol. 20, 781 (1989a).
used by acetogenic bacteria that grow on carbon monoxide or 28. G. Rao and R. Mutharasan, Altered electron flow in con-
carbon dioxide and hydrogen. Proc. Biochem. Sym. 2 9 - 5 6 tinuous cultures of Clostridium acetobutylicum induced by
(1982). viologen dyes. Appl Environ. Microbiol. 53, 1232-1235
10. R. Kerby and J. G. Zeikus, Growth of Clostridium thermo- (June 1987).
aceticum on H2/CO2 or CP as Energy Source. Curr. 29. G. Rao and R. Murtharasan, Altered electron flow in a reduc-
Microbiol. 132, 1 - 9 (1982). ing environment in Clostridium acetobu~licum. Biotechnol.
11. B. R. S. Genthner and M. P. Bryant, Growth of Eubacterium Lett. 10, 129-132 (1988).
limosum with carbon monoxide as the energy source. Appl. 30. D. T. Jones et al., Solvent production and morphological
Environ. Microbiol. 43, 7 0 - 7 4 (1982). changes in Clostridium acetobutylicum induced by viologen
12. F. Mayer, R. Lurz and S. Schoberth, Electron microscopic dyes. Appl Environ. Microbiol. 43, 1434- 1439 (1982).
investigation of hydrogen-oxidizing acetate-forming anaerobic 31. S. Long et al., Isolation of solvent production, clostridial stage
bacterium Acetobacterium woodii. Arch. Microbiol. 115, and endospore formation in Clostridium acetobutylicum. Eur.
207-214 (1977). J. Appl. Microbiol. Biotechnol. 20, 256-261 (1984).
13. R. Sleat, A. Mah and R. Robinson, Acetoanaerobium noterae: 32. J. C. Gottschal and J. G. Morris, Biotechnol. Lett. 3,
new genus, new species, an anaerobic bacterium that forms 525-530 (1981).
acetate from hydrogen and carbon dioxide. Int. J. Syst. 33. C. G. Pheil and Z. G. Ordal, Sporulation of the thermophilic
Bacteriol. 35, 10-15 (1985). anaerobes. Appl. Microbiol. 51, 893-989 (1967).
14. W. E. Balch, S. Schoberth, R. S. Tanner and R. S. Wolfe, 34. J. L. Vega, E. C. Clausen and J. L. Gaddy, Study of gaseous
Acetobacterium, new genus of hydrogen-oxidizing, carbon- substrate fermentations: carbon monoxide to acetate. 1. Batch
dioxide-reducing anaerobic bacteria. Int. J. Syst. Bacteriol. culture. Biotechnol. Bioengng 34, 774-784 (1989b).
27, 355-361 (1977). 35. J. L. Vega, G. M. Antorrena, E. C. Clausen and J. L. Gad@,
15. R. L. Uffen, Anaerobic growth of a Rhodopseudomonas Study of gaseous substrate fermentations: carbon monoxide to
species in the dark with carbon monoxide as sole carbon and acetate. 2. Continuous culture. Biotechnol. Bioengng
energy substrate. Proc. Nam. Acad. Sci. U.S.A. 73, 34, 785-793 (1989c).
3298-3302 (1976). 36. J. E. Bailey and D. F. Ollis, Biochemical Engineering Fun-
16. M. P. Dashekvicz and R. L. Uffen, Identification of a carbon damentas. McGraw-Hill, New York (1986).
monoxide-metabolizing bacterium as a strain of Rhodo- 37. R. K. Finn, Agitation-aeration in the laboratory and in
pseudomonas gelatinosa. Int. J. Systematic Bacteriol. 29, industry. Bacteriol. Rev. 18, 154-274 (1954).
145-148 (1979). 38. J. L. Vega, E. C. Clausen and J. L. Gaddy, Biofilm reactors
17. R. S. Breed, E. G. D. Murray, and N. R. Smith, Bergey's for ethanol production. Enzyme Microb. Technol. 10, 403
Manual of Determinative Bacteriology. The Williams and (1988).
Wilkins Company, Baltimore, MD (8th edn) (1987).

You might also like