Vous êtes sur la page 1sur 8

Clean Techn Environ Policy (2008) 10:211218

DOI 10.1007/s10098-007-0141-4

ORIGINAL PAPER

Siloxane removal from biogas by biofiltration: biodegradation


studies
Francesca Accettola Georg M. Guebitz
Rainer Schoeftner

Received: 10 September 2006 / Accepted: 3 February 2007 / Published online: 22 January 2008
! Springer-Verlag 2008

Abstract Recently a lot of attention has been focused


on volatile methyl siloxanes (VMSs) in biogas because of
the costly problems deriving from the formation of silicate-based deposits in biogas-fuelled power plant
equipments. Currently, VMSs are removed from biogas
with high operational costs by adsorption on activated
carbons. Biofiltration could be a cost-effective and environmentally friendly alternative to current technologies,
leading to a decrease in the cost of biogas treatment,
therefore enhancing its use for power generation. This
document presents the results of biodegradation studies on
VMSs aimed to investigate the possibility of using biofiltration to treat biogas. Growth of bacteria isolated from
activated sludge with octamethylcyclotetrasiloxane (D4)
in the vapour phase as the only carbon source was
observed. 16S-rDNA-sequencing showed that the mixed
population mainly contained c-proteobacteria; within
these, Pseudomonas was the dominating genus. A biotrickling filter was set up to treat different air flows with a
siloxane concentration varying between 45 and 77 mg m-3.
Measurements on the gas output revealed removal efficiency up to 20% compared to a control in sterile
conditions.

F. Accettola (&) ! R. Schoeftner


PROFACTOR Produktionsforschungs GmbH,
Energie-& Umwelttechnologie, Im Stadtgut A2,
4407 Steyr/Gleink, Austria
e-mail: francesca.accettola@profactor.at
G. M. Guebitz
Department of Environmental Biotechnology,
Graz University of Technology, Petersgasse 12,
8010 Graz, Austria

Keywords Siloxane ! Biodegradation ! Biogas !


Pseudomonas

Introduction
Over the past few years, there has been increasing interest
towards biogas for energy production in cogeneration
plants, internal combustion engines and recently also in
fuel cells (Spiegel and Preston 2003, 2000; Spiegel et al.
1999). Its use for energy production is justified by the high
concentration of methane (4070 vol%) and encouraged by
several regulations aiming at the reduction of emissions.
Biogas is mostly produced by the digestion of organic
materials in waste disposal sites and sewage treatment
plants. Besides methane, this gas contains carbon dioxide
(3050 vol%), a lower percentage of N2, H2, CO and a
small amount of contaminants, responsible for damaging
equipments and requiring service interruptions and
expensive repairs. Specific contaminants to biogas utilization are hydrogen sulphide, halides and silicon-containing
compounds (volatile methyl siloxanesVMSs).
The concentration of VMSs in biogas (Schweigkofler and
Niessner 2001) depends on the source of biogas; landfill gas
is usually richer in siloxanes than fermentation gas and the
concentration can be up to 50 mg m-3. One of the most
common compounds is octamethylcyclotetrasiloxane (D4)
shown in Fig. 1. VMSs originate from hydrolysis of polydimethylsiloxane (PDMS) (Chandra 1997), an organosilicon
compound (Fig. 1) used in a wide range of consumer and
industrial applications, released into the environment
through landfills and wastewater treatment plants.
During combustion of biogas, siloxanes oxidize to SiO2.
These crystalline deposits can ultimately build a surface
thickness of several millimetres and they are difficult to

123

212

F. Accettola et al.

Fig. 1 Polydimethylsiloxane
(PDMS) hydrolyzes to lower
molecular-weight compounds
(VMSs) like
octamethylcyclotetrasiloxane
(D4)

remove by chemical or mechanical means. They accumulate on pistons, cylinder heads and valves reducing
compression and engine efficiency. In a 1 MW engine
fuelled with a biogas flow rate of 240 Nm3 h-1 containing
only D4 with a concentration of 12 mg m-3 (1 ppmv),
silicates will generate up to 21 kg silicates per year, part of
which are deposited on the surfaces (Fig. 2) (Applied Filter
Technology 2003). The manufacturers of gas engines
introduced a limit value for silicon of 1 mg l-1 measured
in the oil of gas engines in order to prevent premature
engine failure due to silicon-induced damages (Prabucki
et al. 2001). Other limits imposed by engine manufacturers
refer to the content of methane. In this case the maximum
total siloxane concentration allowed might go down to less
than 10 mg Nm-3 methane (Environment Agency 2002)
(corresponding approximately to 5 mg Nm-3 biogas).
Silicate deposits can also result in poor heat transfer in heat
exchangers. In turbines, silicates can cause abrasions to the
blades. When using biogas in microturbines the limit
imposed is less than 10 ppb (Commonwealth Energy
Biogas 2004). Additionally, the glassy residues are
responsible for the inactivity of the catalyst of the emission
control system. The deposits clog the catalyst bed reducing
the availability of sites where the catalytic reaction occurs,
thus reducing the removal efficiency of combustion

Fig. 2 Silicate deposits on a gas engine

123

products (CO and NOx). Catalyst beds can fail in a few


days and there is no way to recover them. A similar process
occurs when biogas is used in fuel cells (AMONCO 2004):
siloxane contaminations reduce the availability of active
sites of the catalyst, resulting in a lower voltage and power
density for given value of the current density.
Treatment systems currently adopted and reported by
industry sources are typically based on adsorption on
activated carbon or other media (Prabucki 2001; Environment Agency 2002; Commonwealth Energy Biogas 2004;
Stoddart et al. 1999; Digester GasFuel Cell 2002).
Adsorption can be integrated with chilling to decrease the
humidity of the gas, thus enhancing the adsorption of
VMSs. The GEW RheinEnergy fuel cell plant in Cologne
has been operating the first fuel cell plant in Europe, producing energy from the digester gas of the wastewater
treatment plant of the town (Digester GasFuel Cell 2002).
The fuel cell produces 200 kW of electricity, covering the
50% of the energy demand of the plant, while the rejected
heat is used to maintain the optimal temperature in the
fermentation tank. The cleaning device adopted in the plant
consists of a two-stage basic cleaning unit; the process gas
is chilled at a temperature of -30"C and finally purified by
activated carbon. The system can treat 110 m3 h-1 gas,
guaranteeing a purity level of \1 mg siloxane m-3. The
cost of the treatment unit is $107,000 while the maintenance costs, for a working period of 10 years, are estimated
at $16,300 per year. Other reported removal systems are
based on the use of polymorphous graphite and silica gel.
The major inconveniences of these technologies are related
to the costs; the power load required to cool down the gas,
as well as the need to regenerate or to dispose the media,
result in high operational and maintenance costs. Moreover, the adsorption process is non-selective; other
compounds can also adsorb, reducing siloxanes removal
effectiveness very quickly.
The aim of the present work was to investigate the
possibility of the application of a biofiltration system to
eliminate siloxanes from biogas; biological methods for
gas treatment have been successfully implemented for the
removal of hydrogen sulphide in odour control, and their
applicability to hydrogen sulphide removal from biogas is
very promising (AMONCO 2004; U.S. Environmental

Siloxane removal from biogas by biofiltration

Protection Agency 2003; Trogish and Baaske 2004).


Nevertheless, as far as concerning siloxanes, very little is
known about their biodegradability: biodegradation tests
are hard to perform because of their low solubility and high
vaporization rate. For a long time organosiloxanes have
been considered inert with respect to living organism; then
the possibility of their chemical degradation and biodegradation in the environment was proven (Lukasiak et al.
2002). PDMS was found to be biodegraded in soil (Lehmann and Griessbach 1999; Lehmann et al. 1995) through
a two-step mechanism. In the first step re-arrangement and
hydrolysis of PDMS occur with the production of lower
molecular weight compounds (primarily D4 and D5 and
linear VMSs) and silanols (mostly dimethylsilanediol
DMSD). Upon this partial hydrolysis of PDMS to DMSD,
the second step of degradation may proceed, giving inorganic water-soluble compounds as final product. Higher
rates have been detected for the first step when decreasing
the water content, thus leading to the general idea that this
step is purely chemical and catalyzed by clays minerals.
The second step, degradation of DMSD, on the opposite,
should be biological (Lehmann et al. 1998; Stevens 1998).
Sabourin et al. (1996) identified a bacterium, Arthrobacter
sp., and a fungus, Fusarium oxysporum using DMSD as cosubstrate (degradation rates: 2.4% per month for F. oxysporum when adding 2-propanol; 4% per month for
Arthrobacter in presence of dimethylsulfon). Research
conducted by Wasserbauer et al. (1990) demonstrated that
bacteria from the genus Pseudomonas (P. putida and P.
fluorescens) are capable of growing on oligomeric and
polymeric dimethylsiloxanes, with the growth being proportional to the viscosity (hence of the chain length) of
siloxanes. Biodegradation occurs in aerobic conditions.
Degradation products in this study were not identified.
Grumping et al. (1999) found DMSD in sewage sludge
spiked with D4 and incubated in anaerobic conditions. D4
degradation after 100 days was found to be 3% of the
initial added quantity. PDMS degradation was not detected.
Pseudomonas aeruginosa, Proteus mirabilis and Klebsiella pneumoniae, were found to be able to degrade
organosilicone compounds (Rosciszweewski et al. 1998),
with the biodegradation being higher for polyethoxysiloxane than for PDMS. The same research also underlined the
effect of molecular structure on biodegradability: branched
or cross-linked polymers are easier to degrade than longchain structured siloxanes.
Other studies (Slavicek et al. 2000; Varaprath et al.
1999) reported colonization and assimilation of silicone in
voice prosthesis and VMSs metabolism, thus proving
siloxane degradability. Recently, the potential of enzymes
to catalyze the formation and cleavage of siloxane bonds
has been investigated (Brandstadt Var Kurt 2005).

213

The research aimed to isolating bacteria from sludge


growing on siloxanes and studying a biotrickling filter
system, inoculated with the isolated bacteria, treating a
siloxanes-enriched air stream.

Materials and methods


Batch cultures
For biodegradation tests, batch cultures were prepared by
using a mineral medium (Table 1) at pH 7 (series 1). The
only carbon source added was octamethylcyclotetrasiloxane D4 (98%) from Sigma-Aldrich (4 ml of D4 in a
volume of 150 ml mineral medium). Furthermore, in a
second series (series 2) of cultivation experiments, siloxane was not added directly into the aqueous solution but
in a separate container inside the flask; in this way, only
siloxane dissolved in the liquid from the vapour phase
was actually available to the bacteria. This method was
adopted to overcome the difficulty of measuring optical
density in a two-phase sample. Cultures were inoculated
respectively with a pure strain of P. putida DSMZ 291
(DSMZ, Braunschweig, Germany), activated sludge from
the municipal waste-water treatment plant (WWTP) of the
city of Steyr (Austria) and activated sludge from the
WWTP of a company producing silicones. Sludge from a
silicon manufacturer was considered because of the possibility to have microorganisms more adapted to siloxane
degradation. The cultures were periodically re-inoculated
into new medium. A blank culture (without bacteria) and
a zero culture (without D4) were also analysed. The
cultures were grown aerobically in 250 ml Erlenmeyer
flasks filled to 100 ml and shaken on a rotary shaker at

Table 1 Composition of the cultivation medium


Medium composition per liter
K2HPO4

12.5 g

KH2PO4

3.8 g

(NH4)2SO4
MgSO4!7H2O

1g
0.1 g

5 ml of a trace element solution (composition per liter)


H3BO3
ZnSO4!7H2O

FeSO4(NH4)2SO4!6H2O

0.232 g
0.174 g
0.116 g

CoSO4!7H2O

0.096 g

CuSO4!5H2O

8 mg

(NH4)6Mo7O24!4H2O
MnSO4!4H2O

0.022 g
8 mg

123

214

150 rpm at ambient temperature. The growth was monitored by optical density measurements at k = 270 nm
[wavelength related to nucleic acids (Alupoaei and Garcia-Rubio 2004)] and k = 600 nm using a Lambda 35
UV/VIS spectrophotometer (Perkin-Elmer, Austria) and
by measuring protein concentration with a protein test
according to the Bradford method, calibrating the method
by using bovine serum albumin (BSA) as standard. One
millilitre of homogenised sample (by using an Ultra
Turrax IKA T18 at 10,000 rpm for 2 min) was mixed
with 0.5 ml of a 0.3 N NaOH solution and heated up to
60"C for 90 min. Eventually, the samples were centrifuged and the supernatant was used for the determination
of the protein concentration. A protein assay kit was
purchased from Fluka (Austria). Optical density and
protein concentration were measured in duplicated and
homogenised samples.

Analysis of D4 and degradation products


Impurities in D4 have been analysed by means of GCWLD (Agilent GC 6890, Waldbronn, Germany) and
GCMS (Agilent 6890 with LECO TOF-MS Pegasus III).
Degradation products have been detected by H+-NMR
(Bruker, Avance 400) and by performing a silicic acid test

Fig. 3 Biotrickling filter set-up.


The filter has been inoculated by
using batch cultures. The
medium recycled is a mineral
medium

123

F. Accettola et al.

purchased from Dr. Bruno Lange (Germany) on the


supernatant obtained from centrifugation of the samples.
The test is based on the reaction of dissolved silicates with
ammonium molybdate in an acidic solution to form yellowcoloured silicomolybdic acid. If a reducing agent is added a
blue dye is formed and the quantity of silicic acid can be
detected by a spectrophotometer (Lambda 35 UV/VIS,
Perkin-Elmer, Austria).

Microorganism identification
Isolated bacteria have been identified in a first attempt, by
using API 20NE tests, purchased by bioMerieux GmbH,
and in a second attempt by 16S-rDNA-sequencing (analysis performed by Ecowork Laboratories Consulting
GmbH, Vienna, Austria). DNA was isolated from the
provided bacterial culture and used as a template for PCR
amplification of the 16S-rDNA. The obtained producta
mixture of the amplified 16S-rDNA-fragments of the bacteria in the culturewas cloned into plasmid pTZ57R/T;
48 single colonies were picked for RFLP analysis with the
restriction enzyme BsuRI of the cloned insert. Representative inserts from each restriction pattern were sequenced.
After editing, sequences were used for searches against
public databases (BLASTsearch).

Siloxane removal from biogas by biofiltration


Table 2 Gas flow rate, empty
bed retention time and organic
load of the biotrickling filter
with different airflow rates
Organic load remains constant

215

G (m3 h-1)

EBRT (min)

D3 (mg m-3)

Organic load (g m-3 h-1)

0.03 (0.5 l min-1)

3.6

77

1.28

0.04 (0.7 l min-1)

2.7

59

1.28

0.05 (0.9 l min-1)

2.1

46

1.28

Biotrickling filter

Results and discussion

Batch cultures from series 2, showing bacterial growth,


have been used to inoculate a biotrickling filter (Fig. 3). A
biotrickling filter was preferred to a biofilter because of the
necessity to avoid the presence of other carbon sources in
this first step of research. The filter was filled with inert
packing material (Pall Rings 5/800 in plexiglass, free volume 88%, specific surface 350 m2/m3); the height of the
packing material was 0.5 m and the diameter 0.07 m. The
mineral medium was recirculated from the bottom to the
top of the column (by means of a solenoid dosing pump
Prominet gamma/l; flow rate 0.015 m3 h-1) to provide the
bacteria with moisture and minerals necessary for growth.
An air stream containing siloxane was guided through the
column in co-current with the recirculated medium. The
column was covered by aluminium paper to avoid the
formation of algae, which can occur in conditions of low
availability of carbon sources.
Siloxane D3 (Sigma-Aldrich) was used in the experiment
with the biotrickling filter because of the possibility to use a
system (Kin-Tek 491M, Aero-Laser, Garmisch-Partenkirchen, Germany) able to generate a constant concentration of
D3 in the air stream just by adjusting the temperature and the
air flow rate. The system is based on the molecular permeation of vapour through a polymeric membrane, which
establishes a small and stable flow of the compound. This
small flow of permeate is mixed with a controlled flow of
dilution gas to form a precisely known trace concentration in
the gas. With this system, any change in the air flow rate
gives a change of the siloxane concentration, hence the
organic load remains constant (Table 2). The biotrickling
filter inlet tube was insulated to avoid interferences due to
change in temperature. Siloxane D4 was also injected into
the air stream during the experiments, by using a continuousflow syringe pump (TSE System GmbH, Germany) at
2.5 ll h-1, at a concentration of 75 mg m-3.
Siloxane concentration in the gas streams in the biofilter
system was measured by gas chromatograph (Trace GC
2000, ThermoQuest Austin TX, USA) coupled with a flame
ionisation detection. Siloxane was adsorbed on Chromosorb
102 (Supelco, Austria) and extracted by ethyl acetate
(recovery from chromosorb was equal to 99.25%). Octylbenzene was used as internal standard at a concentration of
50 mg l-1. The separation was carried out on a PE-5MS,
0.25 l, 30 m 9 0.32 mm (Perkin Elmer, Austria).

Biodegradation of siloxanes was studied in batch cultures


and biotrickling filters.

Batch cultures
Batch cultures of series 2 showed at first an increase of the
optical density although values were very low (lower than
0.1 for OD600 and 0.4 for OD270). During this first period, it
was not possible to measure the protein content because of
some interference in the test, probably due to contaminants
in the sludge. These low values of OD were measured for a
period of six (for cultures with inoculum from Steyr
WWTP) and nine (inoculum from WWTP of silicon

Fig. 4 Optical density of batch cultures inoculated with activated


sludge. a Activated sludge from a silicon producer WWTP. b
Activated sludge from Steyr WWTP. All values are the average of
duplicate measurements with deviation below 3%

123

216

F. Accettola et al.

degradation seems likely. Previously, biodegradation of


PDMS in soil was described via hydrolysis to lower
molecular weight compounds (Lehmann and Griessbach
1999; Lehmann et al. 1999; Stevens 1998).

Bacteria identification

Fig. 5 Protein concentration in batch cultures

producer) subsequent inoculations into fresh medium,


corresponding to overall cultivation time of 24 and
20 weeks. After this period the culture with P. putida did
not show any particular changes, while the cultures containing bacteria from sludge started to show an evident
increase of the turbidity and high values of optical density
(OD600 up to 2 for WWTP of silicon producer, and 4 for
WWTP Steyr, Fig. 4).
Protein concentration also increased during incubation
to more than 100 lg ml-1 after 70 days of incubation
(Fig. 5). The blank culture (without bacterial inoculum) did
not show any increase of the OD while the zero culture
(without D4) showed increasing values in the first 15 days
(OD600 up to 0.09 and OD270 up to 0.47). However, the OD
dropped to zero after approximately 50 days. In order to
check if the growth was due to the metabolism of impurities contained in the siloxane, the latter was analysed by
means of GCWLD and GCMS. The analysis did not
show other organic solvents in D4 while the major impurities were D3 and D5; hence, growth of bacteria on other
carbon sources may not be assumed.

Degradation products
Based on measurements with the silicic acid test, no significant amounts of silicates were detected in any of the
samples (series 1). It may be possible that silicates were
mostly concentrated in the deposit obtained from the centrifugation. It can also be supposed that silicate interacts
with other intermediate reaction products not detectable as
free silicate or that the degradation does not end with the
formation of dissolved silicates.
The analysis of dimethylsilanediol (DMSD) in the cultures (series 1) revealed a concentration of 14 ppm when
compared to 7 ppm in the blank; therefore, D4 hydrolysis
was favoured by the presence of bacteria. Thus, biotic

123

A first attempt to identify the bacteria in the batch cultures


was made by using API strips 20NE. P. aeruginosa and
Agrobacterium radiobacter were identified as the main
representatives in the mixed population. A following and
more accurate identification was made by 16S-rDNAsequencing. The results revealed that the cultures contained
mainly c-proteobacteria (Pseudomonas citronellosis, P.
putida, Rhodanobacter lindaniclasticus, Zoogloea ramigera, Mesorhizobium sp., Xanthomonadacea) within this
group the genus Pseudomonas was dominating. Previously,
Arthrobacter sp., and a fungus, Fusarium oxysporum have
been described to metabolise DMSD (Sabourin et al. 1996)
while several Pseudomonas species were able to grow on
oligomeric and polymeric dimethylsiloxanes (Wasserbauer
and Zadak 1990; Rosciszweewski et al. 1998).

Biotrickling filter
The removal efficiency of D3 [RE = (Cin - Cout)/Cin] and
elimination capacity [EC = (Cin - Cout) 9 Q/V, where
Cin is the inlet concentration, Cout the outlet concentration,
Q the gas flow and V the empty volume of the filter] have
been evaluated. During these first experiments, D4 was also
injected in the air stream by a continuous-flow syringe
pump; nevertheless, its concentration was not stable,
therefore no data about D4 degradation in the filter are
reported.
The gas flow rate was varied to check its influence in the
removal efficiency. As specified above, an increase in the
flow rate in this system causes a proportional decrease in
the D3 concentration to keep the organic load constant.
Operation of the biotrickling filter was started with a gas
flow rate of 0.5 l min-1 and an inlet D3 concentration of
77 mg m-3. As shown in Fig. 6a, relative high values of
RE and EC occurred when starting the experiment, probably because of some adsorption effect on the packing
material. During the following days, the RE was quite
stable between 10 and 20% while the EC was around
0.2 g m-3 h-1. The peaks in the graphs (days 2, 14, 30,
44) correspond to the substitution of the medium with fresh
one; hence it is possible that the higher values are due to a
major effect of absorption and to better nutritional conditions for the bacteria. After about 35 days though, the
efficiency decreased at values lower than 10%, excluding

Siloxane removal from biogas by biofiltration

Fig. 6 D3 removal efficiency (%) of the biotrickling filter at different


gas flow rates. a gas flow rate = 0.5 l min-1; b gas flow rate = 0.9
l min-1. All values are based on triplicate measurements with a
relative standard deviation between 0.56 and 3.92%

the higher values reported after changing the medium. At


this point, the flow rate of the gas stream was increased to
evaluate its effect. When operating at 0.9 l min-1, which
corresponds to an inlet concentration of D3 of 45 mg m-3,
the removal efficiency was well above 10% (Fig. 6b) while
the EC was around 0.2 g m-3 h-1. During these experiments the medium had been changed only once at day 57.
In a next step, the gas flow rate was decreased to 0.7 l
min-1, which led to a concentration of 59 mg D3 m-3.
From this moment, D4 was no longer injected. The medium was changed only once at day 64. The operation was
quite stable, with an RE around 10%.
After the operational period, the biotrickling filter was
emptied for an estimation of the amount of biomass. The
packing material was washed with distilled water and the
resulting water plus biomass was dried in the oven at
105"C for 48 h. The resulting weight of the dried matter
was about 62 g.
The biotrickling filter operated again after cleaning and
sterilization of the filling material to evaluate the effect of
absorption. The results showed no removal of siloxane in
absence of a biological system. Therefore, D3 removal was
not due to physical absorption on the medium or to
adsorption on the carrier material. The absence of
absorption is, as expected, due to the low mass transfer

217

coefficient (which is connected with the design of the


system) and the relatively high value of the partition
coefficient which is linked to the physical characteristics of
siloxanes (Henrys law constant for D4 = 0.117 atm m3
mole-1), which are not soluble in water and volatile.
Therefore, the transfer of the compound from the gas into
the liquid phase is strongly limited; nevertheless, the
physical properties in a gas-aqueous biomass system can
differ from a gas-pure water system and result in a better
mass transfer. For example, the solubility of some hydrocarbons is higher in aqueous biomass than in pure water
(Barton et al. 1997; Davidson et al. 2000). This is due to
the ability of some bacteria (among them the genus Pseudomonas) to produce bio-surfactants to increase the
availability of a substrate present in a gas phase, thus
leading to better performances of biotrickling filters than
expected when considering a normal air-water system
(Tuleva et al. 2002; Zhang and Miller 1995). The solubility
of D3 could increase because of the presence of biomass,
leading to an improvement of the mass transfer and thus a
major availability of the compound for the bacteria. The
removal reported in the biofilter, despite the low solubility
of the system, may also be explained assuming an
agglomeration of the bacteria at the gas-liquid interface and
therefore a more direct access to the substrate. The system
may be developed by improving the mass transfer (for
example by using surfactants and/or a structured packing
instead of a random packing) and with a better understanding of the kinetics of the microbiological process.

Conclusions
Siloxane biodegradation has been investigated in order to
evaluate the possibility of using a biofiltration system to
treat biogas.
Results show that octamethylcyclotetrasiloxane can be
biodegraded by a community of microorganisms isolated
from activated sludge. Pseudomonas was identified as the
predominant genus in the mixed population while other
microorganisms found included Rhodanobacter, Zooglea,
Mesorhizobium and Xanthomonadacea. Furthermore,
DMSD, the degradation product of D4, has been found in
bacterial cultures in a higher concentration than in the
corresponding blank. A biotrickling filter was set up for the
treatment of an air stream polluted with siloxane. The
results obtained showed a removal of 1020% D3, while
the same system in abiotic conditions reported no removal
at all.
The present work on siloxane biodegradation and further
optimisation of the process (in regards to the biofilter setup and the determination of microbiologic parameters)
would give an important contribution to an issue which, so

123

218

far, has been hardly addressed and about which very few
data are available from literature. The major aim would be
the development of a low-cost unit able to upgrade biogas
by biological means, thus encouraging its use for power
generation. Biofiltration has been proved a viable, low-cost
option for facilities having emissions that qualify for this
technology. It has been applied successfully for odour
control, treatment of air contaminated with hydrocarbons
or other pollutants, and, in regards to H2S, for biogas
upgrading. Its application for siloxane removal would be an
interesting cost-effective and environmentally friendly
alternative to present technologies for avoiding damage to
equipment and increase of CO and NOx emissions in biogas-fuelled power plants.
Acknowledgments This work has been carried out with the financial support of European Community (Marie Curie Host Fellowship).

References
Alupoaei CE, Garcia-Rubio RH (2004) Growth behaviour of microorganisms using UVvis spectroscopy: Escherichia coli.
Biotechnol Bioeng 86(2):163167
AMONCO EU Project (2004) Advanced prediction, monitoring and
controlling of anaerobic digestion behaviour towards biogas
usage in fuel cells. Contract No. ENK6-CT-2001-00518.
http://www.eva.ac.at/projekte/biogas_fuelcell.htm. Accessed 23
September 2007
Applied Filter Technology (2003) New technology for removal of
siloxanes in digester gas results in lower maintenance costs and
air quality benefits in power generation equipment. Paper
presented at WEFTEC 03 78th annual technical exhibition and
conference, 1115 October. http://www.appliedfiltertechnology.
com. Accessed 23 September 2007
Barton JW, Llasson KT, Koran LJ, Davison BH (1997) Microbial
removal of alkanes from dilute gaseous waste streams: kinetics
and mass transfer considerations. Biotechnol Prog 13:814821
Brandstadt Var Kurt F (2005) Inspired by nature: an exploration of
biocatalyzed siloxane term bond formation and cleavage. Curr
Opin Biotechnol 16(4):393397
Chandra G (1997) The handbook of environmental chemistry, part H.
Organosilicon materials. Springer, Heidelberg
Commonwealth Energy Biogas/PV Mini-Grid Renewable Resources
program (2004) Making renewables part of an affordable and
diverse electric system in California. Contract No. 500-00-036,
Project No. 2.2. Enhanced energy recovery through optimization
of anaerobic digestion and microturbines; Task 2.2.1 Final
report. http://www.energy.ca.gov/distgen_oii/documents/dgwg/
R+D-12.pdf. Accessed 23 September 2007
Davidson BH, Barton JW, Klasson KT (2000) Influence of high
biomass concentration on alkane solubilities. Biotechnol Bioeng
68:279284
Digester GasFuel Cell Project (2002) Final report for the U.S.
Department of Energy. http://www.dodfuelcell.com/climate/
reports/41045_Final.pdf. Accessed 23 September 2007
Environment Agency (2002) Guidance on gas treatment technologies
for landfill gas engines. Draft for Consultation, Environment
Agency, UK. http://www.sepa.org.uk/pdf/consultation/closed/
2003/gas/gas_treatment.pdf. Accessed 23 September 2007

123

F. Accettola et al.
Grumping R, Michalke K, Hirner AV, Hensel R (1999) Microbial
degradation of octamethylcyclotetrasiloxane. Appl Environ
Microbiol 65:22762278
Lehmann RG, Griessbach FC (1999) Degradation of polydimethylsiloxane fluids in the environmenta review.
Chemosphere 38(6):14611468
Lehmann RG, Varaprath S, Annelin RB, Arndt JL (1995) Degradation
of silicone polymer in a variety of soils. Environ Toxicol Chem
14:12991305
Lehmann RG, Miller JR, Collins HP (1998) Microbial degradation of
dimethylsilanediol in soil. Water Air Soil Pollut 106(12):11
122
Lukasiak J et al (2002) Biodegradation of silicones (organosiloxanes).
In: Biopolymers, vol 9. Wiley, New York, pp 539568
Prabucki MJ, Doczyck W, Asmus D (2001) Removal of organic
silicon compounds from landfill and sewer gas. In: Proceedings
Sardinia 2001, eighth international waste management and
landfill symposium, Cagliari, Italy, vol 2. Cisa, Cagliari, pp
631639
Rosciszweewski P, Lukasiak J, Dorosz A, Galinski J, Szponar M
(1998) Biodegradation of polyorganosiloxanes. Macromol Symp
130:337346
Sabourin CL, Carpenter JC, Leib TK, Spivack JL (1996) Biodegradation of dimethylsilanediol in soils. Appl Environ Microbiol
62:43524360
Slavicek A, Betka J, Novak V, Kinkal J, Wasserbauer R, Vymazalova
Z, Lastovickova B (2000) Biodegradation of silicone in voice
prosthesis. Sb Lek 101(4):325329
Spiegel RJ, Preston JL (2000) Test results for fuel cell operation on
anaerobic digester gas. J Power Sources 86:283288
Spiegel RJ, Preston JL (2003) Technical assessment of fuel cell
operation on anaerobic digester gas at the Yonkers, NY,
wastewater treatment plant. Waste Manage 23:709717
Spiegel RJ, Preston JL, Trocciola JC (1999) Fuel cell operation on
landfill gas at Penrose power station. Energy 24:723742
Schweigkofler M, Niessner R (2001) Removal of siloxanes in
biogases. J Hazard Mater B83:183196
Stevens C (1998) Environmental degradation pathways for the
breakdown of polydimethylsiloxanes. J Inorg Biochem 69:203
207
Stoddart J, Zhu M, Staines J, Rothery E, Lewicki R (1999)
Experience with halogenated hydrocarbons removal from landfill
gas. In: Proceedings Sardinia 1999; Seventh international waste
management and landfill symposium, Cagliari, Italy, vol 2. Cisa,
Cagliari, pp 489498
Trogish S, Baaske WE (2004) Biogas powered fuel cells. Trauner
Verlag, Linz
Tuleva BK, Ivanov GR, Christova NE (2002) Biosurfactant production by a new Pseudomonas putida strain. Z Naturforsch
57c:356360
U.S. Environmental Protection Agency (2003) Using bioreactors to
control air pollutionU.S. Environmental Protection Agency,
EPA-456/R-03-003. http://www.epa.gov/ttn/catc. Accessed 23
September 2007
Varaprath S, Salyers KL, Plotzke KP, Nanavati S (1999) Identification of metabolites of octamethilcyclotetrasiloxane (D4) in rat
urine. Drug Metab Dispos 27(11):12671273
Wasserbauer R, Zadak Z (1990) Growth of Pseudomonas putida and
P. fluorescens on silicone oils. Folia Microbiol 35:384393
Zhang Y, Miller RM (1995) Effect of rhamnolipid (Biosurfactant)
structure on solubilization and biodegradation of n-alkanes. Appl
Environ Microbiol 61(6):22472251

Vous aimerez peut-être aussi