You are on page 1of 5

International Biodeterioration & Biodegradation 52 (2003) 1 – 5

www.elsevier.com/locate/ibiod

Degradation of organic pollutants by the white rot basidiomycete


Trametes trogii
L. Levina; ∗ , A. Vialeb , A. Forchiassina
a MicologaExperimental, Departamento de Cs. Biologicas y, Facultad de Ciencias Exactas Naturales, Universidad de Buenos Aires,
1428 Capital Federal, Argentina
b Microbiologa, Departamento de Qumica Biol
ogica, Facultad de Ciencias Exactas Naturales, Universidad de Buenos Aires,
1428 Capital Federal, Argentina

Abstract

The ability of the white rot basidiomycete Trametes trogii (strain BAFC 463) to degrade in vitro concentrations of 250 –500 ppm of
nitrobenzene and anthracene was analyzed. Within 12 – 24 days, more than 90% of the organic pollutants added to the fungal cultures were
removed, as demonstrated by gas–liquid chromatography. Enzyme estimations indicated a high and relatively stable activity of laccase, as
well as a lower production of manganese peroxidase. Laccase activity could be implicated in the degradation of these xenobiotic compounds
by T. trogii. Earlier work showed that this fungus almost completely removed anthraquinone dyes and polychlorinated biphenyls. In view
of the results obtained, this strain seems promising for detoxi5cation. ? 2002 Elsevier Science Ltd. All rights reserved.

1. Introduction (Beaudette et al., 1998), dioxins (Joshi and Gold, 2000),


pesticides (Kullman and Matsumura, 1996), explosives
Polycyclic aromatic hydrocarbons (PAHs) are pollu- (Hawari et al., 1999) and dyes (Abadulla et al., 2000).
tants, often carcinogenic, mutagenic, or toxic, found in Intracellular systems that are generally present in most
most terrestrial ecosystems, that arise from industrial oper- fungi, such as cytochrome P-450 monooxygenase, may
ations and from natural events such as forest 5res. Major also be involved in organopollutant degradation (Bezalel
components of petroleum, they are continuously released et al., 1996). However, as ligninolytic enzymes are active
into natural environments, posing a serious risk to human extracellularly, white rot fungi are better candidates for the
health (Sack et al., 1997). Bioremediation of PAH-polluted bioremediation of highly apolar pollutants than nonligni-
soil is severely hampered by the low rate of degradation nolytic microorganisms (Field et al., 1993).
of higher PAHs, particularly the four- and 5ve-ring PAHs. BTEX compounds (benzene, toluene, ethylbenzene, and
These higher PAHs have very low water solubility and are o-, m- and p-xylenes) are an important family of organic
often tightly bound to soil particles. This results in very low pollutants that are components of gasoline and aviation
bioavailability for bacterial degradation. The observation fuels; they and their nitro derivatives are widely used in
that white rot fungi can oxidize PAHs rapidly with their ex- industrial syntheses. They are carcinogenic and neurotoxic
tracellular ligninolytic enzyme systems has therefore raised and are classi5ed as priority pollutants (Spain, 1995). Both
interest in the use of these organisms for bioremediation aerobic and anaerobic bacteria have been shown to degrade
of PAH-polluted soils (Kotterman et al., 1998). Oxidation BTEX compounds, but most of these studies on bacterial
of PAHs by white rot fungi to more water-soluble prod- degradation of BTEX have used microbial consortia and
ucts with greater bioavailability resulted in higher rates of no native strain of bacterium is known to degrade all the
mineralization by bacteria than those of the parent com- components of BTEX eFciently. O-xylene particularly has
pounds (Meulenberg et al., 1997). The nonspeci5c oxidases been shown to be recalcitrant to bacterial degradation (Tsao
involved in lignin degradation (mainly laccases, lignin- et al., 1998). On the other hand, the white rot fungus Phane-
and manganese-peroxidases) are able to transform PAHs rochaete chrysosporium simultaneously degraded all the
(Johannes and Majcherczyk, 2000), chlorinated phenols BTEX components (Yadav and Reddy, 1993; Song, 1997).
(Reddy and Gold, 2000), polychlorinated biphenyls (PCBs) Most of the research on bioremediation by white rot fungi
has focused on this single species which is known to metab-
∗ Corresponding author. olize a wide range of xenobiotic compounds (Paszczynski
E-mail address: aaviale@qb.fcen.uba.ar (L. Levin). and Crawford, 1995). Recently, however, other fungi have

0964-8305/03/$ - see front matter ? 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 9 6 4 - 8 3 0 5 ( 0 2 ) 0 0 0 9 1 - 4
2 L. Levin et al. / International Biodeterioration & Biodegradation 52 (2003) 1 – 5

begun to be evaluated for their pollutant-degrading abil- MnSO4 and H2 O2 and 0:01% phenol red as the substrate.
ities, and notable diJerences with regard to the extent of Reactions were halted by adding NaOH 5 N, and increase
their pollutant transformation ability have been demon- in A610 measured (E610 = 22 mM−1 cm−1 ). Lignin peroxi-
strated. The utilization of fungi for bioremediation requires dase activity (LiP) was assayed by the method of Tien and
an understanding of the factors that enhance their ability to Kirk (1988) monitoring the increase in absorbance at 310
detoxify, and clari5cation of the enzyme mechanisms used nm due to oxidation of veratryl alcohol to veratryl aldehyde
(Cerniglia et al., 1992). (E310 = 9:333 mM−1 cm−1 ). One unit of enzyme activity
Trametes trogii is a white rot basidiomycete, distributed was de5ned as the amount of enzyme required to oxidize
worldwide. T. trogii strain BAFC 463, besides eFciently 1  mol of substrate min−1 . Enzyme activity is expressed as
degrading lignin in wood (Levin and Castro, 1998), has been U ml−1 of culture 5ltrate. The extracellular proteins were
tested successfully in biomechanical pulping experiments measured by the Bradford method (Bradford, 1976) with
(Planes et al., 1986) and also demonstrated to be a good pro- bovine serum albumin as the standard. The results are the
ducer of ligninases (Levin and Forchiassin, 2001). The si- average of three triplicate experiments with a standard error
multaneous presence in this fungus of high ligninolytic and of ¡ 5%.
hydrogen-peroxide-producing activities, essential for per- Degradation of xenobiotics. The degradation of an-
oxidase activity and rate limiting for pollutant degradation thracene (3-ring PAH) and nitrobenzene by the fungus was
(Gramss et al., 1999), make it an attractive microorgan- measured by adding 250 –500 ppm to the 25 ml Erlenmeyer
ism on which to base future biotechnological applications. cultures in their 4th day of growth. After a total incuba-
Previously it was shown to almost completely remove an- tion period of either 12 or 24 days, adding HClO4 to pH
thraquinone dyes and PCBs (Haglund, 1999; Levin et al., 1 stopped the growth. The remaining compounds were ex-
2001). In this study its ability to transform nitrobenzene and tracted either with methylene chloride (for nitrobenzene) or
anthracene in liquid cultures was analyzed in relation to its benzene (for anthracene) and analyzed by gas chromatog-
complement of extracellular ligninolytic enzymes. raphy (GC) coupled to a mass spectrometer TRIO 2, in
a capillar column (50  m, 0.32 ID) with helium as car-
rier gas at 0.7 psi, temperature program: 150◦ C for 1 min,
2. Materials and methods temperature increase 5◦ C min−1 and 240◦ C, for 10 min.
Uninoculated medium controls and heat-killed culture con-
Microorganism. Strain 463 (BAFC: Mycological Culture trols were also included. The identi5cation of the diJerent
Collection of the Department of Biological Sciences, Faculty chemicals was carried out by analyzing the fragmentation
of Exact and Natural Sciences, University of Buenos Aires) pro5les with the aid of Mass Lynx v 2.1. This software al-
of T. trogii Berk. (Aphyllophorales, Basidiomycetes) was lows the quantitative analysis of the kinetics of degradation
used. Stock cultures were maintained on malt extract agar by integrating the area under each peak, which is a direct
slants at 4◦ C. measure of the concentration of the chemical.
Basal culture medium. Was GA (Levin and Forchiassin,
2001) with glucose 10 g l−1 and asparagine 3 g l−1 . The 5nal
pH of the medium was adjusted to 4.5 with 0.5 M citric acid 3. Results and discussion
and 0.5 M dibasic sodium phosphate buJer, diluted in the
medium to a 5nal concentration of 0.25 M. The kinetics of The kinetics of in vitro production of extracellular ligni-
growth and enzyme production was studied in this medium nolytic activities by T. trogii were studied in a synthetic
and also in a medium containing glucose (20 g l−1 ) and malt medium (glucose 10 g l−1 =asparagine 3 g l−1 ) and in a com-
extract (20 g l−1 ). plex medium (malt extract 20 g l−1 =glucose 20 g l−1 ). High
Culture conditions. Erlenmeyer Masks of 250 ml contain- activities of laccase and MnP were detected in both media
ing 25 ml of medium were inoculated with 2 agar plugs (each (but especially in the synthetic medium), laccase activity
of 0:25 cm2 ) cut out from the margin of a colony grown on being predominant. Both activities appeared before mycelial
Bacto-agar 2%. Incubation was carried out at 28 ± 1◦ C un- biomass peaked (primary phase of growth: trophophasic
der stationary conditions. Cultures were harvested after dif- cultures), but their highest levels were detected during the
ferent incubation periods, 5ltered through 5lter paper using phase of secondary metabolism (idiophasic growth). Maxi-
a BNuchner funnel and dried overnight at 70◦ C. Dry weight mal levels detected were: 0:55 U ml−1 laccase, 0:07 U ml−1
of mycelia was then determined. The culture supernatants MnP in malt extract=glucose medium; 4:1 U ml−1 laccase,
were used as enzyme sources. 0:22 U ml−1 MnP in glucose=asparagine (Fig. 1). Attempts
Enzyme assays. Laccase activity was determined by mea- to detect LiP in the culture media were unsuccessful. The
suring the increase in A420 due to the oxidation of 0.5 mM negative LiP tests suggests that the fungus produces no
ABTS (2; 2 -azinobis(3-ethylbenzthiazoline-6-sulphonic signi5cant levels of this enzyme or its production requires
acid) in 0.1 M sodium acetate buJer (pH 5.0) (E420 = diJerent growth conditions. LiP activity was detected pre-
36 mM−1 cm−1 (Bourbonnais et al., 1995). Manganese viously, when this strain was grown in a wood-containing
peroxidase activity (MnP) was assayed with 0.1 mM medium (Levin and Forchiassin, 2001).
L. Levin et al. / International Biodeterioration & Biodegradation 52 (2003) 1 – 5 3

4.2 240

MnP [mU/ml]/ Dry weight [mg/25ml]


3.6 210

180
3.0
Laccase [U/m l]

150
2.4
120
1.8
90
1.2
60
0.6 30

0.0 0
0 3 6 9 12 15 18 21 24 27
Culture age [days]

Fig. 1. Kinetics of growth and enzyme production by Trametes trogii


in a synthetic medium (glucose 10 g l−1 =asparagine 3 g l−1 ) (A) and in
a complex medium (malt extract 20 g l−1 =glucose 20 g l−1 ) (B). ()
Growth (mycelial dry weight (mg=25 ml)) in (A), () growth in (B),
( ) laccase activity (U=ml) in (A), () laccase activity (U=ml) in (B),

() MnP activity (mU=ml) in (A), ( ) MnP activity in (B).

Fig. 3. Consume of 250 and 500 ppm of anthracene by trophophasic


cultures of T. trogii grown in glucose=asparagine medium: (A) abiotic
control, (B) 500 ppm, (C) 250 ppm. The remaining hydrocarbon was
extracted with benzene.

cultures of T. trogii grown in glucose=asparagine medium


in the presence of 500 ppm of nitrobenzene. Only 3% of
the nitrobenzene was recovered from trophophasic cultures
(Fig. 2B; abiotic control Fig. 2A: 100%) and none from id-
iophasic cultures (Fig. 2C).
On Fig. 3 the consumption of 250 and 500 ppm of an-
thracene by trophophasic cultures of T. trogii grown in
glucose=asparagine medium is shown. Either 10% or 5%,
of respectively, 500 and 250 ppm of anthracene, added to
viable cultures, were recovered at trophophasic phase (Fig.
3B and C). Similar results were obtained in the complex
medium.
The resolved GC peaks were identi5ed and matched with
the fragmentation pro5les in the MassLynx spectral library
(Fig. 4). Chromatograms after 12 days of incubation in malt
extract=glucose medium with anthracene showed the pres-
Fig. 2. Chromatograms of methylene chloride extracts obtained from cul- ence of a peak (retention time 10.07 min) (Fig. 4B), absent
tures of T. trogii grown in glucose=asparagine medium in the presence of in the control incubated without anthracene (Fig. 4A). This
500 ppm of nitrobenzene: (A) abiotic control (12 days), (B) trophophasic peak was identi5ed as 4-methoxybenzaldehyde.
culture (12 days), (C) idiophasic culture (24 days). Most of the white rot fungi tested to date (Bjerkandera,
Phanerochaete and others) produce quinone intermediates
T. trogii showed considerable tolerance to the xenobi- during the degradation of PAHs, but it was demonstrated
otic compounds assayed; concentrations of up to 500 ppm previously that strains of the genus Trametes degraded an-
anthracene and 250 ppm nitrobenzene did not signi5cantly thracene without any accumulation of anthraquinone (Field
aJect its growth (data not shown). With 500 ppm of ni- et al., 1992). Identical results were obtained in the present
trobenzene, mycelial dry weight decreased 50%. study.
Fig. 2 shows the chromatograms obtained from the ex- A number of nonligninolytic fungi have intracellular
traction of trophophasic (12 days) and idiophasic (24 days) mechanisms for PAH degradation that usually lead to
4 L. Levin et al. / International Biodeterioration & Biodegradation 52 (2003) 1 – 5

addition of known precursors and aromatic acids represent-


ing lignin degradation products stimulated the production
of aryl metabolites (Mester et al., 1997). Similarly the pres-
ence of anthracene or its degradation products could induce
the synthesis of the aryl compound detected in our study.
T. trogii in contrast to the well studied model organism
P. chrysosporium, is N-unregulated. SuFcient or excess
N-nutrients stimulate high MnP and laccase titres in parallel
with the high biomass production. This characteristic of the
fungus makes it an outstanding candidate for large-scale
fermentation to produce ligninolytic enzymes in bulk for
bioremediation. It produces high amounts of MnP, and
higher laccase levels than those reported for most other
white rot fungi under favorable conditions, accompanied
by high levels of the hydrogen-peroxide-producing enzyme
glyoxal oxidase (Levin and Forchiassin, 2001). The pres-
ence of extracellular peroxidases, but also of laccase, may
be crucial to an eJective oxidation of PAH adsorbed to soil
substrates. An eJective conversion of PAH in biotechno-
logical soil remediations may nevertheless greatly depend
on the presence of H2 O2 in the substrate or its generation
by the fungus itself (Gramss et al., 1999). Although the
Fig. 4. Mass spectra of the metabolite detected after incubation of T. trogii role of individual ligninolytic enzymes in PAH-degradation
in the presence of anthracene. In the corner: chromatograms of benzene
is still not completely understood, it was shown recently
extracts obtained from T. trogii grown in malt extract=glucose medium
(A) for 12 days without anthracene, to which 250 ppm of anthracene were that both laccase and MnP are capable or PAH degra-
added immediately before the extraction, (B) 250 ppm of anthracene were dation in the presence of appropriate mediators secreted
added to the cultures at 4th day of growth, and extracted 8 days later. by ligninolytic fungi (Gramss et al., 1999; Johannes and
Majcherczyk, 2000). The BTEX-degradation system is ex-
pressed during primary metabolism in P. chrysosporium
dihydrodiol and hydroxy PAH metabolites. These prod- (Yadav and Reddy, 1993; Song, 1997). Extracellular lignin-
ucts are suspected to originate from the activity of cy- and manganese-peroxidase are not produced during this
tochrome P-450 monooxygenases in conjunction with stage; probably laccase or other hitherto uncharacterized
epoxide hydrolase. Epoxides and dihydrodiols are very po- enzyme(s) may be involved in their oxidation. In view of
tent carcinogens. Even P. chrysosporium has been shown the results currently being obtained in our laboratory on the
to metabolize phenanthrene under nonligninolytic condi- biodegradation of diJerent organopollutants by T. trogii,
tions to dihydrodiol and hydroxylated PAH metabolites this strain seems promising for detoxi5cation.
(Cerniglia et al., 1992). Such compounds were not detected
in this study. As suggested by Bezalel et al. (1996), it is
probable that both ligninolytic and cytochrome P-450 en- Acknowledgements
zymes are involved in the degradation of PAHs in vivo.
Cytochrome P-450 enzymes perform the initial modi5ca- To CONICET and University of Buenos Aires for 5nan-
tion of the aromatic groups followed by degradation of cial support.
the PAH derivative by the ligninolytic system. The pres-
ence of both a PAH-degrading cytochrome P-450 and a
ligninolytic system in P. chrysosporium supports this hy- References
pothesis (Masaphy et al., 1996). Aryl metabolites, such as
4-methoxybenzaldehyde, which are produced simultane- Abadulla, E., Tzanov, T., Costa, S., Robra, K.-H., Cavaco, P.A., Gubitz,
G.M., 2000. Decolorization and detoxi5cation of textile dyes with a
ously with the extracellular ligninolytic enzymes, have an laccase from Trametes hirsuta. Applied Environmental Microbiology
important role in the ligninolytic system of white rot fungi. 66, 3357–3362.
The most commonly occuring aryl compounds are veratryl Beaudette, L.A., Davies, S., Fedorak, P.M., Ward, O.P., Pickard,
(3,4 -dimethoxybenzyl) and p-anisyl (4-methoxybenzyl) M.A., 1998. Comparison of gas chromatography and mineralization
alcohol–aldehyde. Veratryl alcohol is known to be a co- experiments for measuring loss of selected polychlorinated biphenyl
congeners in cultures of white rot fungi. Applied Environmental
factor involved in the degradation of lignin, lignin model
Microbiology 64, 2020–2025.
compounds and xenobiotic pollutants by LiP. P-anisyl al- Bezalel, L., Hadar, Y., Cerniglia, C.E., 1996. Mineralization of polycyclic
cohol is thought to be involved in H2 O2 generation as a aromatic hydrocarbons by the white rot fungus Pleurotus ostreatus.
substrate for the extracellular aryl alcohol oxidase. The Applied Environmental Microbiology 62, 292–295.
L. Levin et al. / International Biodeterioration & Biodegradation 52 (2003) 1 – 5 5

Bourbonnais, R., Paice, M.G., Reid, I.D., Lanthier, P., Yaguchi, M., 1995. Levin, L., Forchiassin, F., 2001. Ligninolytic enzymes of the white rot
Lignin oxidation by laccase isozymes from Trametes versicolor and basidiomycete Trametes trogii. Acta Biotechnologica 21, 179–186.
role of the mediator 2,2-azinobis (3-ethylbenzthiazoline-6-sulphonate) Levin, L., Jordan, A., Forchiassin, F., Viale, A., 2001. DegradaciQon
in kraft lignin depolymerization. Applied Environmental Microbiology de azul de antraquinona por Trametes trogii. Revista Argentina de
61, 1876–1880. Microbiologia 33, 223–228.
Bradford, M.M., 1976. A rapid and sensitive method for the quantitation Masaphy, S., Levanon, D., Henis, Y., Venkateswarlu, K., Kelly,
of microgram quantities of protein utilizing the principle of protein-dye S.L., 1996. Evidence for cytochrome P-450 and P-450-mediated
binding. Analytical Biochemistry 72, 248–254. benzo(a)pyrene hydroxylation in the white rot fungus Phanerochaete
Cerniglia, C.E., Sutherland, J.B., Crow, S.A., 1992. Fungal metabolism chrysosporium. FEMS Microbiology Letters 135, 51–55.
of aromatic hydrocarbons. In: Winkelmann, G. (Ed.), Microbial Mester, T., Swarts, H.J., Romero SolQe, S., de Bont, J.A.M., Field, J.A.,
Degradation of Natural Products. VCH Press, Weinheim, pp. 193–217. 1997. Stimulation of aryl metabolite production in the basidiomycete
Field, J.A., de Jong, E., Feijoo Costa, G., de Bont, J.A.M., 1992. Bjerkandera sp. strain BOS55 with biosynthetic precursors and lignin
Biodegradation of polycyclic aromatic hydrocarbons by new isolates of degradation products. Applied Environmental Microbiology 63, 1987–
white rot fungi. Applied Environmental Microbiology 58, 2219–2226. 1994.
Field, J.A., de Jong, E., Feijoo Costa, G., de Bont, J.A.M., 1993. Screening Meulenberg, R., Rijnaarts, H.H.M., Doddema, H.J., Field, J.A., 1997.
for ligninolytic fungi applicable to the biodegradation of xenobiotics. Partially oxidized polycyclic aromatic hydrocarbons show an increased
Trends in Biotechnology 11, 44–49. bioavailability and biodegradability. FEMS Microbiology Letters 152,
Gramss, G., Kirsche, B., Voight, K.-D., Gunther, T., Fritsche, W., 1999. 45–49.
Conversion rates of 5ve polycyclic aromatic hydrocarbons in liquid Paszczynski, A., Crawford, R.L., 1995. Potential for bioremediation
cultures of 5fty-eight fungi and the concomitant production of oxidative of xenobiotic compounds by the white-rot fungus Phanerochaete
enzymes. Mycological Research 103, 1009–1018. chrysosporium. Biotechnology Progress 11, 368–379.
Haglund, C., 1999. Biodegradation of xenobiotic compounds by the white Planes, E., Bassi, M., Bensignor, J., Burachik, M., 1986. Pulpado biolQogico
rot fungus Trametes trogii. Master’s degree project, Upsala University de maderas de sauce. ATIPCA 25, 44–46.
School of Engineering. Reddy, G.J.B., Gold, M.H., 2000. Degradation of pentachlorophenol by
Hawari, J., Halasz, A., Beaudet, S., Paquet, L., Ampleman, G., Phanerochaete chrysosporium: intermediates and reactions involved.
Thiboutot, S., 1999. Biotransformation of 2,4,6-trinitrotoluene with Microbiology 146, 405–413.
Phanerochaete chrysosporium in agitated cultures at pH 4.5. Applied Sack, U., Heinze, T.M., Deck, J., Cerniglia, C.E., Martens, R., Zadrazil,
Environmental Microbiology 65, 2977–2986. F., Fritsche, W., 1997. Comparison of phenanthrene and pyrene
Johannes, C., Majcherczyk, A., 2000. Natural mediators in the oxidation degradation by diJerent wood-decaying fungi. Applied Environmental
of polycyclic aromatic hydrocarbons by laccase mediators systems. Microbiology 63, 3919–3925.
Applied Environmental Microbiology 66, 524–528. Song, H.-G., 1997. Biodegradation of aromatic hydrocarbons by several
Joshi, D.K., Gold, M.H., 2000. Oxidation of dibenzo-p-dioxin by lignin white-rot fungi. Journal of Microbiology 31, 66–71.
peroxidase from the basidiomycete Phanerochaete chrysosporium. Spain, J.C., 1995. Biodegradation of nitroaromatic compounds. Annual
Biochemistry 33, 10969–10976. Review of Microbiology 49, 523–555.
Kotterman, M.J.J., Vis, E.H., Field, J.A., 1998. Successive mineralization Tien, M., Kirk, T.K., 1988. Lignin peroxidase of Phanerochaete
and detoxi5cation of benzo(a)pyrene by the white rot fungus chrysosporium. Methods in Enzymology 161B, 238–249.
Bjerkandera sp. strain BOS55 and indigenous microMora. Applied Tsao, C.-W., Song, H.-G., Bartha, R., 1998. Metabolism of benzene,
Environmental Microbiology 64, 2853–2858. toluene, and xylene hydrocarbons in soil. Applied Environmental
Kullman, S.W., Matsumura, F., 1996. Metabolic pathways utilized by Microbiology 64, 4924–4929.
Phanerochaete chrysosporium for degradation of the cyclodiene Yadav, J.S., Reddy, C.A., 1993. Degradation of benzene, toluene,
pesticide endosulfan. Applied Environmental Microbiology 62, 593– ethylbenzene, and xylenes (BTEX) by the lignin-degrading
600. basidiomycete Phanerochaete chrysosporium. Applied Environmental
Levin, L., Castro, M.A., 1998. Anatomical study of the decay caused by Microbiology 59, 756–762.
the white-rot fungus Trametes trogii (Aphyllophorales) in wood of
Salix and Populus. IAWA Journal 19, 169–180.

You might also like