ArticlePDF Available

White-rot fungi-mediated bioremediation as a sustainable method for xenobiotic degradation

Authors:

Abstract and Figures

Xenobiotics are hazardous compounds that are foreign to natural ecosystems. The production and use of xenobiotic compounds continue to increase worldwide, which in turn causes environmental pollution and has adverse effects on humans. Degradation of such compounds, therefore, needs urgent awareness and attention. The physicochemical approaches to treat the contaminants are expensive. Bioremediation is a concept that exploits organisms to manage the environment with less manpower and time. White-rot fungi-mediated bioremediation offers inexpensive, environmentally sustainable and potential degradation mechanisms for different recalcitrant chemicals. White-rot fungi secrete lignolytic enzymes that have extensive substrate specificities and are involved in the transformation and solubilization of lignin-like structural contaminants. The main lignolytic enzymes occurring in white-rot fungi are laccases, lignin peroxidase, manganese peroxidase, and other peroxidases. Such lignolytic enzymes permit white-rot fungi to endure high toxic levels. This review describes the opportunities to use white-rot fungi and their enzyme systems in the biodegradation of multiple xenobiotic contaminants.
Content may be subject to copyright.
103
Introduction
Xenobiotics are toxic chemicals that are exotic to living
organisms and have an anity to persist in the biosphere
(Sinha et al. 2009). ese compounds have synthetic
chemical composition, and species have not adapted to
these in evolution (Gren 2012). e residues of xenobiotic
substances persist in the ecosystem over a long period and
have negative eects on the microora and the fertility of
the soil (Gianfreda, Rao 2008). erefore, polluting the
environment with recalcitrant chemicals, which oen
are xenobiotics, is one of the major environmental issues
with global focus and recognition (Tišma et al. 2010).
e pollution created by xenobiotics disrupts natural
ecosystems, causes changes in climatic conditions, reduces
water levels and has other negative impacts (Gursahani,
Gupta 2011). e main sources of xenobiotics enter into the
environment from pharmaceutical industries (ibuprofen,
paracetamol), agriculture (pesticides, herbicides,
insecticides), the paper industry (paper and pulp euent),
food industry (food additives such as vinegar, lecithin),
plastic industry (polyvinyl chloride), consumer industry
(coatings, dyes) and petroleum industries (benzene, xylene)
(Mishra et al. 2019). Humans are exposed to xenobiotics
through inhalation, adsorption by skin (cosmetic
products) or ingestion (medicines, vegetables, fruits).
ey can cause severe health hazards such as heart defects,
neurodegeneration, defects in the central nervous system
and adverse reproductive problems. Hence, xenobiotic
degradation in the environment is essential (Phale et al.
2019).
Physico-chemical approaches involved in the
control of organic pollutants include ion exchange,
chemical occulation, adsorption, irradiation, oxidation,
precipitation and ozonation (Aksu 2005). e physico-
chemical approaches are very costly and oen yield adverse
intermediate metabolites that are harmful and need further
secondary treatment. To overcome these, several other
environmentally friendly processes have been described,
such as bioremediation, phytoremediation, etc. (Varsha
et al. 2011). Bioremediation is a technology in which
biological organisms (algae, bacteria, fungi and plants) are
employed to minimize the accumulation and harmfulness
of environmental contaminants (Gnanasalomi et al. 2013).
Xenobiotic microbial depletion is an eective strategy
for eliminating toxic pollutants from the environment.
Environmental and Experimental Biology (2021) 19: 103–119 Review
http://doi.org/10.22364/eeb.19.11
White-rot fungi-mediated bioremediation
as a sustainable method for xenobiotic
degradation
A. Kathiravan, J. Joel Gnanadoss*
Microbial and Environmental Biotechnology Research Unit, Department of Plant Biology and
Biotechnology, Loyola College, Chennai-600034, Tamil Nadu, India
*Corresponding author, E-mail: joelgna@gmail.com
ISSN 2255-9582
Abstract
Xenobiotics are hazardous compounds that are foreign to natural ecosystems. e production and use of xenobiotic compounds
continue to increase worldwide, which in turn causes environmental pollution and has adverse eects on humans. Degradation of
such compounds, therefore, needs urgent awareness and attention. e physicochemical approaches to treat the contaminants are
expensive. Bioremediation is a concept that exploits organisms to manage the environment with less manpower and time. White-rot
fungi-mediated bioremediation oers inexpensive, environmentally sustainable and potential degradation mechanisms for dierent
recalcitrant chemicals. White-rot fungi secrete lignolytic enzymes that have extensive substrate specicities and are involved in the
transformation and solubilization of lignin-like structural contaminants. e main lignolytic enzymes occurring in white-rot fungi are
laccases, lignin peroxidase, manganese peroxidase, and other peroxidases. Such lignolytic enzymes permit white-rot fungi to endure
high toxic levels. is review describes the opportunities to use white-rot fungi and their enzyme systems in the biodegradation of
multiple xenobiotic contaminants.
Key words: biodegradation, lignolytic enzymes, mycoremediation, white-rot fungi, xenobiotics.
Abbreviations: BTEX, benzene, toluene, ethylbenzene and xylene; DyP, dye-decolourizing peroxidase; LiP, lignin peroxidase; MnP,
manganese peroxidase; PAHs, polyaromatic hydrocarbons; PCBs, polychlorinated biphenyls; PCPs, pentachlorophenols; TNT,
2,4,6-trinitrotoluene; YMG, yeast-malt-glucose
104
e ability of microorganisms to break down xenobiotic
substances is considered an important means of removing
toxic materials (Sridevi et al. 2011). Some fungi are resilient
organisms compared to bacteria and are typically less
susceptible to high levels of pollutants. is is why fungi
were extensively studied regarding their bioremediation
capabilities in the mid-1980s (Ellouze, Sayadi 2016).
Fungi play a crucial role in all environments including
soil and marine habitats as decomposers and symbionts.
ey are especially suitable for bioremediation because
of their robust morphological structure and various
biochemical capabilities. Mycoremediation is a component
of bioremediation that employs fungus for intrinsic and
extrinsic management of polluted areas. Mycoremediation
is a cost-eective and environmentally reliable option
for removing, transporting and storing hazardous waste.
Mycelia may destroy these contaminants within the soil
before they move through food chains (Ramachandran,
Gnanadoss 2013). Further, attention has been given to the
distinct ability of fungi to remove these contaminants by
using a variety of extracellular and intracellular enzyme
systems for detoxication and bioremediation (Deshmukh
et al. 2016). e objective of this review paper is to
emphasize the signicant properties of white-rot fungi in
degrading dierent xenobiotic compounds like polymers,
polyaromatic hydrocarbons (PAHs), polychlorinated
biphenyls (PCBs), pentachlorophenols (PCPs),
antibiotics, 2,4,6-trinitrotoluene (TNT), benzene, toluene,
ethylbenzene and xylene (BTEX), pesticides and dyes.
White-rot fungi
White-rot fungi possess the remarkable ability to
biodegrade lignin and hence the name white-rot comes
from the white surface of timber invaded by white-
rot fungi, where the evacuation of lignin gives a faded
impression (Pointing 2001). Systematically, white-rot fungi
include specic basidiomycetes, and very few ascomycete
families like Xylariaceae are associated with white-rot decay
(Eaton, Hale 1993). e utilization of fungi for the cleaning
of contaminated soil was initially demonstrated during
the mid-1980s, when the white-rot fungus Phanerochaete
chrysosporium was found to degrade a high diversity of
natural contaminants (Bumpus, Aust 1987). is ability
was later exhibited for diverse species such as Trametes
versicolor and Pleurotus ostreatus (Ghani et al. 1996),
Lentinus subnudus (Adenipekun, Fasidi 2005), Psathyrella
candolleana LCJ 178 and Myrothecium gramineum
LCJ 177 (Gnanasalomi, Gnanadoss 2013), Porostereum
spadiceum (Tigini et al. 2013), Pleurotus floridanus LCJ155,
Leucocoprinus cretaceous LCJ164, and Agaricus sp. LCJ169
(Jebapriya, Gnanadoss 2014), Dentipellis sp. (Park et al.
2019), Ganoderma lucidium (Coelho-Moreira et al. 2018)
and Bjerkandera adusta (Dhiman et al. 2020). White-rot
fungi degrade all timber components, for example, cellulose,
hemicellulose and lignin, whereas other fungi destroy
lignin predominantly. e former is called a non-selective
white-rot degraders and the latter are known as specic
white-rot degraders. e specic white-rot degraders are
extremely intriguing from the biotechnology perspective as
they remove lignin leaving the lucrative cellulose unaltered
(Dashtban et al. 2010). Potential benets of using white-rot
fungi to remove ecological contaminants are due to their
ubiquitous existence, ability to break down assorted classes
of destructive foreign substances and to adjust the pH of
their characteristic substrate (Christian et al. 2005).
Enzyme systems in white-rot fungi
White-rot fungi usually produce one or multiple lignolytic
enzymes in various amounts, on the basis of which they
can be divided into four classes (Nerud, Misurcova
1996) namely: (a) laccase, manganese peroxidase (MnP)
and lignin peroxidase (LiP), (b) laccase and any of the
peroxidases, (c) laccases only, (d) peroxidases only. e
most widely recognized lignolytic enzymes present in
white-rot fungi incorporate laccases and MnP and the
least common are LiP and versatile peroxidase. ese
lignolytic enzymes can work together or independently,
yet additional enzymes like glyoxal oxidase, aryl alcohol
oxidase, cellobiose dehydrogenase, pyranose 2-oxidase,
and others are fundamental to accomplish the cycle of
lignocellulose or xenobiotic degradation. In addition, an
intracellular enzyme cytochrome P450 monooxygenase
and low molecular weight oxidants like hydroxyl radicals
and Mn3+ were demonstrated to be powerful in eliminating
lignocellulosic materials and various xenobiotics. Recently,
dye decolourising peroxidase (DyP), which is involved in
the decolouration of dyestus, and aromatic peroxygenases
has been found to be associated in catalysis of oxygen
transfer reactions that bring about the ester cleavage, is also
recognized as a lignolytic enzyme that corresponds with
white-rot fungi (Rodríguez-Couto, 2016).
Laccase
Laccase is a copper protein that has its position in blue
oxidases. Copper, which occurs at the dynamic site of
the enzyme, plays an integral role in catalytic reactions.
e catalytic center of the enzyme consists of four copper
atoms. Laccase catalyzes four single electron oxidations
of the substrate into four electron reductive bond
cleavages. Degradation of various aromatic mixtures
can be catalyzed by an associative reduction of oxygen
to water. Moreover, in the presence of key substrates
[2,20-azinobis-3-ethylbenzothiazoline-6-sulphonicacid
or 1-hydroxybenzotriazole] working as electron transfer
mediators, the substrate spectrum is further extended to
degrade non-phenolic mixtures (Kılıç et al. 2016). Laccase
was rst recognized in 1883 by Yoshida, when he separated
the exudates from Rhus vernicifera (urston 1994).
A. Kathiravan, J.J. Gnanadoss
105
ey are derived from natural sources and oen occur
in plants and microorganisms (Dwivedi et al. 2011) and
also in a few insects (Xu 1999). Fungal sources of laccase
have been isolated from dierent groups of fungi like
deuteromycetous, ascomycetous and basidiomycetous. Of
these, white-rot fungi and other litter degrading organisms
are the most prominent sources of the laccase enzyme. In
specic, laccase production by basidiomycetous taxa such as
Trametes, Pleurotus, Agaricus, Phanerochaete, Pycnoporus,
and Lentinus has been broadly explored as they are easy to
grow in in vitro (Rodríguez-Couto, 2019). White-rot fungal
species that synthesize laccase are Polyporus sanguineus,
Phlebia brevispora, Daedalea flavida and Phlebia radiata
(Arora, Gill 2001), Phanerochaete chrysosporium, Trametes
hirsuta, Marasmius sp. and Trametes versicolor (Risdianto et
al. 2012), Pleurotus florida, Pleurotus ostreatus and Pleurotus
sajor-caju (Radhika et al. 2013), Agaricus sp. LCJ262 (Jose,
Joel 2014), Trametes orientalis (Zheng et al. 2017), Cerrena
unicolor strain GSM-01 (Wang et al. 2017), and Myrothecium
gramineum LCJ 177 (Gnanasalomi, Gnanadoss 2019).
Laccases from white-rot fungi are associated with lignin
removal and are resilient at dierent pH and temperatures.
High purity of laccase can be obtained by suitable
optimizing parameters (Gnanasalomi, Gnanadoss 2013).
Laccase-mediator methods have enormous potential for
lignin removal, biosensor application, biofuel and organic
synthesis, bioremediation of some toxic chemical wastes,
pharmaceutical and nanobiotechnology applications
(Singh, Gupta 2020).
Lignin peroxidase
LiP is a heme enzyme from the oxidoreductase family, which
is primarily secreted by white-rot basidiomycetes during
the formation of secondary metabolites. LiP plays a key part
in removing the lignin portion of the plant cell wall. LiP
assists in the biodegradation of lignin and other phenolic
molecules with H2O2 as a substrate and veratryl alcohol
as a mediator (Singh et al. 2019). LiP has been reported in
various white-rot fungi like Coriolus versicolor f. antarcticus
(Levin et al. 2004), Phanerochaete chrysosporium (Wang
et al. 2008), Ganoderma lucidium (Sasidhara et al. 2014),
Pleurotus ostreatus, Pleurotus sapidus, Pleurotus florida
(Kunjadia et al. 2016), Porodaedalea pini (Tanabe et al.
2016), Podoscypha elegans (Agarwal et al. 2017), Coriolopsis
gallica, Pleurotus sajor-caju and Lentinula edodes (Ding et
al. 2019). LiP is exploited for numerous industrial uses and
bioremediation processes due to its immense substrate
specicity and high redox potential (Erden et al. 2009).
Manganese peroxidase
As LiP, MnP is also placed under the same family of
oxidoreductases described in Phanerochaete chrysoporium
as another lignolytic enzyme (Paszczyńskib et al. 1985).
MnP seems to more prevalent in white-rot fungi than
LiP (Hammel, Cullen 2008). In contrast to LiP, MnP has
a low redox potential and oxidizes the compounds with
H2O2 performing as oxidant and manganese performing
as a mediator in the MnP catalytic process. e function
of MnP is the conversion of Mn2+ ions to Mn3+. Mn3+ is
extremely reactive and chelates with biomolecules such as
oxalate and malate formed by the fungus (Shanmugapriya
et al. 2019). Chelated Mn3+ stimulates the degradation
of phenolic compounds to phenoxy radicals (Hofrichter
2002). A few examples of MnP from white-rot fungi are
Bjerkandera sp. (Mester, Field 1997), Irpex flavus, Polyporus
sanguineus and Dichomitus squalens (Gill, Arora 2003),
Physisporinus rivulosus (Hakala et al. 2005), Pleurotus
ostreatus, Coriolus versicolor and Phlebia tremellosa
(Robinson et al. 2011), Cerrena unicolor (Zhang et al. 2018)
and Pseudolagarobasidium sp. (amvithayakorn et al.
2019). MnP nds wide applications in the industries such
as food, textile, paper and pulp industries, pharmaceutical
and bioremediation (Singh et al. 2019).
Versatile peroxidase
Versatile peroxidase is a hybrid peroxidase that comprises
the catalytic activities of MnP and LiP (Dosoretz, Reddy
2007). Similar to MnP, it have a strong anity for Mn2+
and initiates the conversion of Mn2+ to Mn3+ and it also
metabolizes non-phenolic and phenolic molecules without
Mn2+ like LiP. is enzyme seems to be expressed mostly in
fungal genera such as Pleurotus, Bjerkandera, and Lepista
and may also be present in Panus and Trametes (Yadav,
Yadav 2015). Versatile peroxidase has distinctive characters
compared to other lignolytic peroxidases and is suitable
for utilization in various applications like the paper and
pulp industry, biofuel production, ruminant nutrition,
bioremediation, and the textile industry (Ravichandran,
Sridhar 2016).
Dye-decolourising peroxidase
DyP is a novel group of heme peroxidases that were
molecularly identied and are well-known in bacteria and
fungi. ey lack structural and sequence resemblances
with traditional ora and fauna peroxidases. As the name
species, DyP can use H2O2 to detoxify dierent groups
of azo and anthraquinone-based articial dyes, substrates
that are less susceptible to oxidation by members of
other classical heme peroxidases. Also, specic DyP were
described to oxidize compounds of the phenol lignin type,
thus providing the enzymatic ability for this category
of heme peroxidase to support the transformation of
lignocellulosic materials for downstream production of
biofuel (Chaplin et al. 2019). DyP was initially identied
from Bjerkandera adusta culture (formerly reported as
Geotrichum candidum) (Fernández-Fueyo et al. 2015).
DyP producing white-rot fungi include Termitomyces
albuminosus (Johjima et al. 2003), Pleurotus ostreatus
(Faraco et al. 2007), Marasmius scorodonius (Pühse et
al. 2009), Auricularia auricula-judae (Liers et al. 2010),
White-rot fungi mediated bioremediation for xenobiotic degradation
106
Exidia glandulosa, Mycena epipterygia (Liers et al. 2013),
Irpex lacteus (Salvachúa et al. 2013), Funalia trogii (Kolwek
et al. 2018), Trametes versicolor (Amara et al. 2018), and
Pleurotus sapidus (Krahe et al. 2020).
Cytochrome P450 monooxygenase
Cytochrome P450 monooxygenase is a main intracellular
enzyme that ts the class of oxygenases that helps in the
degradation of xenobiotics via oxygen. ey also have
heme-comprising enzymes that incorporate one or several
oxygen molecules to break down aromatic rings and even
stabilize the compound (Baker et al. 2019). e importance
of cytochrome P450 monooxygenase mechanisms in
detoxication of endogenous and exogenous molecules has
been shown (Ichinose et al. 2013). Increased removal of
PAHs was achieved by the initial application of cytochrome
P450 monooxygenase in degradation experiments.
Improved elimination of contaminants was obtained
through molecular methods for ecient and oversupply of
cytochrome P450 enzyme (Deshmukh et al. 2016).
Xenobiotic degradation process by white-rot fungi
e capability of white-rot fungal species in eliminating
xenobiotic compounds from ecosystems is dependent
on their ability to biodegrade lignin, as it is close to the
structure of dierent xenobiotics (Fig. 1). erefore, the
identical methodologies that provide white-rot fungi the
potential to degrade lignin are being used to remove a large
range of xenobiotic contaminants. Most of the xenobiotics
are oxidized and mineralized to various sizes using white-
rot fungi under the lignolytic conditions (Field et al. 1993).
e xenobiotic degradation process is shown in Fig. 2.
Fig . 1. Structure of lignin in comparison with the chemical
structure of diverse xenobiotic compounds.
Fig . 2. Schematic diagram representing the degradation mechanism of xenobiotic compounds using white-rot fungi.
A. Kathiravan, J.J. Gnanadoss
107
Applications of white-rot fungi in xenobiotic
degradation
Degradation of polymers using white-rot fungi
Plastic is typically a polymer that is composed of various
elements such as carbon, hydrogen, silicon, oxygen, chloride
and nitrogen (Seymour 1989). Linking the monomers
using chemical bonds forms plastic. Polythene contains
about 64% of plastic, a continuous hydrocarbon polymer
comprising elongated strands of ethylene monomers.
e overall equation for polyethylene is CnH2n, where n,
represents the total number of carbon molecules (Sangale et
al. 2012). e worldwide usage of plastic is rising rapidly at
a rate of 12% per year and approximately 0.15 billion units
of synthetic materials are developed globally every year
(Das, Kumar 2014). Every year, the ecosystem accumulates
25 million metric tons of polymer waste (Kaseem et al.
2012).
Biodegradation of many synthetic polymers with
dierent chemical compositions has been reported, but
several of them involved degradation using white-rot
fungi-mediated lignin enzymes (Pointing 2001). Nylon-6
polymer degradation initially using white-rot fungus
Bjerkandera adusta has been described (Friedrich et al.
2007). Synthetic materials including polyvinyl chloride,
nylon, acrylamide, etc. that are degraded by dierent
white-rot fungi have been documented (Kale et al. 2015).
Removal of biopolymers like lignin, hemicellulose and
cellulose is also possible by white-rot fungi. In comparison
to other groups of microorganisms, lignin removal through
white-rot fungi is more promising (Woiciechowski et al.
2013). Phellinus pini, Phanerochaete chrysosporium, Phlebia
sp., Pleurotus sp., Heterobasidion annosum, Ceriporiopsis
subvermispora, Irpex lacteus and Trametes veriscolor are
specic white-rot fungi that preferably invade lignin more
eectively than cellulose and hemicellulose. ey secrete
dierent classes of lignolytic enzymes that facilitate the
degradation of aromatic organic compounds, producing
aromatic radicals, and modify the structure of lignin-
and lignocellulose-derived products (Andlar et al. 2018).
Study on the oxidation of the biopolymer lignin from the
paper industry was conducted to determine the capacity
for degradation by ve white-rot fungi (Lentinus edodes,
Pleurotus ostreatus, Trametes versicolor, Phanerochaete
chrysosporium and S22). Among these ve isolates, three
white-rot fungi (Phanerochaete chrysosporium, Pleurotus
ostreatus and S22) showed a high level of lignin degradation
at pH 9.0 to 11.0 (Wu et al. 2005).
Degradation of PAHs using white-rot fungi
PAHs are organic substances that mostly lack colour or
are pale yellow solids. ey are an omnipresent class
of many chemically related compounds that persist in
the environment with complex structures and toxicity
(Abdel Shafy et al. 2016). Some white-rot fungi including
BKM-F-1767, Bjerkandera adusta CBS 595.78, Trametes
versicolor Paprican 52, Phanerochaete chrysosporium and
Trametes versicolor has been tested for ability to degrade
hydrocarbons (Field et al. 1992). All white-rot fungi
signicantly degraded anthracene, as well as nine of the
strains eectively degraded benzo(a)pyrene. Of those,
Bjerkandera sp. Bos 55 seems to be new species and was
deemed an eective degrader of anthracene (99.2 %)
and benzo(a)pyrene (83.1%) molecules within 28 days
respectively. e genus Phanerochaete and Bjerkandera
transformed anthracene into anthraquinone, which is an
end metabolite. Further, this analysis showed Trametes sp.
degraded anthracene with no substantial accumulation of
quinone (Field et al. 1992).
e removal of PAHs by means of white-rot fungi
is aected by temperature, the composition of the
medium, dissolved oxygen and soil moisture content
(Chen et al. 2005). e biological removal of PAHs such
as phenanthrene, uorine and pyrene was achieved using
thermotolerant Trametes polyzona RYNF13. is fungus
exhibited PAH degradation at 100 mg L–1. Complete
removal of phenanthrene was detected in mineral salt
glucose medium at 30 °C aer an incubation period of
18 days while 52% of pyrene and 90% of uorine could
be removed under similar conditions. is fungus is still
capable of surviving at a high temperature of about 42
°C and degrades phenanthrene (68%), uorine (48%)
and pyrene (30%) respectively within 32 days. us, this
strain has potential for PAH degradation specically in
the tropical area where even air temperature can be more
than 40 °C (Teerapatsakul et al. 2016). It was shown that
the most ecient laccase-producing white-rot fungi
Pycnoporus sanguineus can remove phenanthrene (45.6%)
and benz(a)anthracene (90.1%) in in vivo conditions (Li
et al. 2018). ey also transformed phenanthrene into
2-dibenzofuranol by the cytochrome P450 monooxygenase
enzyme or 9,10-phenanthrenedione through extracellular
laccase and benz(a)anthracene into benz(a)anthracene-
7,12-dione through extracellular laccase. Various PAHs
that are degraded by diverse white-rot fungi are represented
in Table 1.
Degradation of PCBs using white-rot fungi
PCBs are widespread organic molecules that were used as
coolant liquids in transformers and electric motors during
the 20th century (Borja et al. 2005). Although their application
and production were prohibited in the last decade, they
survive in ecosystems and lead to severe consequences for
living organisms (Colvin, Nelson 1990). Eaton (1985) was
the rst researcher to study PCBs (Aroclor® 1254 mixture)
degradation using Phanerochaete chrysosporium. In this
experiment, Aroclor® 1254 was mineralized into CO2 with
the removal of water-soluble organic compounds and
irreversible attachment to cells. e results obtained have
been validated by the absence of gas chromatographic peaks.
White-rot fungi mediated bioremediation for xenobiotic degradation
108
PCB congener degradation by Ceriporia sp. was carried out
by Hong et al. (2012). In this analysis, four PCB congeners
(4,4’-dichlorobiphenyl, 2,2’,4,4’,5,5’-hexachloro-biphenyl,
2,3’,4’,5-tetrachlorobiphenyl and 2,2’,4,5,5’-pentachloro-
biphenyl) were examined. e biodegradation rate of
4,4’-dichlorobiphenyl on the 13th day was reported to
be around 34.03% whereas the biodegradation rate of
2,2’,4,4’,5,5’-hexachlorobiphenyl on the 17th day was nearly
40.05%. is shows that extremely chlorinated biphenyls
can be reduced by Ceriporia sp. e immobilized laccase
obtained from Coprinus comatus on wood biochar from
diverse species can degrade chlorinated biphenyl in
wastewater (Li et al. 2018). Pleurotus ostreatus, Trametes
versicolor, Phlebia brevispora, Pycnoporus cinnabarinus, and
Pleurotus sajor-caju are few other white-rot fungi explored
for PCBs degradation.
Degradation of PCPs using white-rot fungi
PCPs are lethal compounds that extensively occur in the
industrial output of pesticides and wood preservatives
(Czaplicka 2004). e usage of PCPs was forbidden
in several countries owing to their high toxicity in the
late 1980s but still, they are used in a few countries. e
harmfulness of pentachlorophenol has been broadly
reported as a recalcitrant and global pollutant in the soil
and water (Varela et al. 2017). Degradation of PCPs is
Tab le 1. Degradation of various PAHs by dierent white-rot fungi
PAH compounds White-rot fungi Reference
Acenapthene Phanerochaete chrysoporium Bishnoi et al. 2008
Anthracene Pleurotus ostreatus, Coriolopsis polyzona, Phanerochaete
chrysosporium, Trametes versicolor
Vyas et al. 1994
Trametes pocas, Trametes cingulate Tekere et al. 2005
Irpex lacteus Baborová et al. 2006
Anthracophyllum discolor Acevedo et al. 2011
Chrysene Bjerkandera sp. Valentin et al. 2007
Polyporus sp. Hadibarata et al. 2009
Pleurotus ostreatus Nikiforova et al. 2010
Pleurotus sajor-caju Saiu et al. 2018
Dibenzothiophene Coriolopsis gallica Bressler et al. 2000
Bjerkandera sp. Valentin et al. 2007
Agrocybe aegerita, Coprinellus radians Aranda et al. 2009
Fluoranthene Bjerkandera sp. Valentin et al. 2007
Phanerochaete chrysosporium Bishnoi et al. 2008
Pleurotus pulmonarius Wirasnita, Hadibarata 2016
Fluorene Agrocybe sp. Chupungars et al. 2009
Pleurotus eryngii Hadibarata, Kristanti 2014
Polyporus sp. Lazim, Hadibarata 2016
Trametes sp. Zhang et al. 2016
Ganoderma sp. Torres-Farradá et al. 2019
Napthalene Phlebia lindtneri Mori et al. 2003
Trametes versicolor Bautista et al. 2010
Armillaria sp. Hadibarata et al. 2012
Pleurotus ostreatus Sukor et al. 2012
Pleurotus eryngii Hadibarata et al. 2013
Ganoderma sp. Torres-Farradá et al. 2019
Phenanthrene Trametes versicolor Han et al. 2004
Bjerkandera sp..Terrazas-Siles et al. 2005
Phanerochaete chrysoporium Bishnoi et al. 2008
Anthracophyllum discolor Acevedo et al. 2011
Ganoderma lucidum Agarwal et al. 2018
Pycnoporus sanguineus Li et al. 2018
Pyrene Dichomitus squalens, Pleurotus sp..Martenz, Zadrazil 1996
Phlebia brevispora Lee et al. 2016
Coriolopsis brysina Agarwal, Shahi 2017
Pleurotus sajor-caju Saju et al. 2018
Quinoloine Pleurotus ostreatus Zhang et al. 2007
A. Kathiravan, J.J. Gnanadoss
109
achieved in three ways: via oxygenolysis, hydroxylation
or reductive dehalogenation (Field, Sierra-Alvarez 2008).
Fungal remediation of PCPs gained interest in the last forty
years. White-rot fungi can degrade PCPs and transform
the respective PCPs compounds through methylation and
dechlorination reactions. PCPs degradation under both
lignolytic and non-lignolytic conditions using three white-
rot fungi (Trametes sp., Pleurotus sp., and Phanerochaete
chrysosporium) were studied (Ryu et al. 2000). e activity
of lignolytic enzymes was detected in Pleurotus and
Trametes cultures, but not in Phanerochaete chrysosporium.
is proves that PCP degradation can be carried out
in two conditions (lignolytic and non-lignolytic) using
white-rot fungi. Phlebia acanthocystis, a white-rot fungus,
was capable of degrading 100% and 76% of PCPs (25 μM
concentration) in low nitrogen as well as potato dextrose
broth culture media, respectively, during incubation for
approximately 10 days (Xiao, Kondo 2020).
e reduction of PCPs in Phlebia acanthocystis
culture is followed by the production of two metabolites
(p-tetrachlorohydroquinone and pentachloroanisole)
through oxidative metabolism. e metabolism of both
molecules is closely linked to Phlebia acanthocystis
extracellular enzymes. Further, the breakdown of PCPs to
p-tetrachlorohydroquinone is carried out by cytochrome
P450 monooxygenase (Xiao, Kondo 2020). e white-rot
fungi with ability to degrade PCPs are Trametes versicolor
(Walter et al. 2004), Anthracophyllum discolor (Rubilar et al.
2007), Bjerkandera adusta, Fomes fomentarius, Ganoderma
applantum, Pleurotus ostreatus, and Laetiporus cincinnatus
(Ramesh, Pattar 2009) and Phlebia acanthocystis (Xiao,
Kondo 2020).
Degradation of TNT using white-rot fungi
TNT is oen utilized as an explosive by the military and
can cause pollution to soil and water at TNT production
and storage sites. It is mutagenic and harmful to many
organisms in that environment. Based on animal studies,
TNT can be a carcinogen for humans (Honeycutt et al.
1996). Most of the white-rot fungi can convert TNT to
dinitrotoluenes and further, result in mineralization to CO2
(Pointing 2001). TNT degradation (50 mg L–1) was studied
using seven white-rot species in two dierent media: yeast-
malt-glucose (YMG) medium and nutrient-rich YMG
medium (Kim, Song 2000). e degradation rate was
higher in nutrient-rich YMG medium than in the limited
nutrient YMG medium. Hydroxylamino-dinitrotoluene
isomers have been recognized as the rst TNT metabolites
to be detected and these compounds are converted into
amino-dinitrotoluenes during further incubation. It was
observed that TNT (90 mg L–1) was degraded in nutrient
broth during 21 days by four white-rot fungal species:
Phanerochaete chrysosporium (67%), Phanerochaete sordida
(87%), Phlebia brevispora (90%), and Cyathus stercoreus
(94%). e TNT degradation from culture was evaluated
by high-performance liquid chromatography and the
cytotoxicity of pollutants in the medium was calculated by
Salmonella/microsome bioassay. is study showed that
white-rot fungi can degrade and detoxify TNT compounds
in aerobic conditions in non-lignolytic nutrient broth
(Donnelly et al. 1997). TNT degradation was examined by
several other white-rot fungi like Irpex lacteus (Kim, Song
2003), Hypholoma fasciculare (Perkins et al. 2005), Trametes
versicolor (Cheong et al. 2006), Kuehneromyces mutabilis,
and Stropharia sp. (Serrano-González et al. 2018).
Degradation of BTEX compounds using white-rot fungi
BTEX compounds like benzene, toluene, ethylbenzene,
and xylenes are a vital class of organic contaminants that
are constituents of petroleum fuels and they are oen
used in various industries as industrial solvents (Smith
1990). e rst study of white-rot fungi degradation
of BTEX compounds was reported by Yaddav, Reddy
(1993). Waste mushroom biomass from Ganoderma
lucidum and Pleurotus ostreatus was used as a substrate
(10%) to decontaminate Esfahan Oil Renery’s petroleum
hydrocarbon contaminated soil (Mohammadi-Sichani et
al. 2019). Petroleum hydrocarbons at the contamination
site were oxidized by waste mushroom biomass of Pleurotus
ostreatus (69.5%) as well as Ganoderma lucidum (57.7%)
and also reduced the soil toxicity in 3rd month respectively.
BTEX compound degradation using white-rot fungi has
been less explored.
Degradation of antibiotics using white-rot fungi
Antibiotics are substances that help to treat communicable
diseases in animals, humans, cattle and aquacultures around
the globe. e discharge of a high proportion of vaccines
into water sources and soil creates a potential threat to all
microbes in these surroundings (Cycoń et al. 2019). e
production of antibiotics is continuously increasing and
their usage has extended from 100 000 to 200 000 tons
worldwide (Gelband et al. 2015). Almost all antibiotics are
not entirely processed in humans and animal bodies. A large
number of therapeutic drugs are discharged into soil and
water bodies by community wastewater, livestock manure,
sludge from sewage and nitrogenous wastes that are oen
used to irrigate and enhance farmlands (Bouki et al. 2013).
Usually, the traditional treatment process is not eective in
treating many pharmaceutical products (Heberer 2002).
e white-rot fungi-mediated bioremediation technique
is therefore a simple and economical method to eliminate
antibiotics.
An in vitro study was conducted on use of LiP obtained
from Phanerochaete chrysosporium for the removal of two
drugs (carbamazepine and diclofenac) that are commonly
found in water bodies. It showed that the LiP entirely
removed diclofenac at pH 3.0 to 4.5 and 3 to 24 ppm H2O2.
e ecacy of carbamazepine degradation is generally
below 10% (Zhang, Geißen 2010). A study on ciprooxacin
degradation using Pleurotus ostreatus (Singh et al. 2017)
showed that the highest degradation rate occurred at
White-rot fungi mediated bioremediation for xenobiotic degradation
110
a concentration of 500 ppm due to maximum enzyme
production. e degradation rate of ciprooxacin aer
14 days of culture at this concentration was evaluated by
three assays: titrimetric (68.8%), indigo carmine (94.25%)
and methyl orange (91.34%). Elimination of ciprooxacin
was additionally demonstrated by high-performance liquid
chromatography, which showed 95.07% degradation and the
microbiological experiment exhibited reduced biological
activities of degraded products against pathogenic bacteria
i.e. Staphylococcus aureus, Escherichia coli and Streptococcus
pyogenes. Trametes hirsuta eciently degraded higher
concentrations of chloramphenicol (10 mg L–1) in the
existence of a laccase mediator system like syringaldehyde,
vanillin, 2,20-azinobis-3-ethylbenzothiazoline-6-sulpho-
nic acid and α-naphthol. e availability of mediators
enhanced the degradation percentage from 10 to 100%
during 48 h. Liquid chromatography mass spectrometry
conrmed the chloramphenicol degradation. e produc-
tion of chloramphenicol aldehyde aer the breakdown
was non-pathogenic to microorganisms (Navada, Kulal
2019). Magnetic cross-linked enzyme aggregates of
Cerrena laccase have been demonstrated to be eective
in the biodegradation of antibiotics such as tetracycline,
oxytetracycline, ampicillin, sulfamethoxazole and erythro-
mycin (Yang et al. 2017). For example, at 40 U mL–1 Cerrena
laccase removed tetracycline antibiotic (100 μg mL–1) at
pH 6 and temperature at 25 °C during 48 h without redox
mediators. Numerous studies revealed that white-rot fungi
have the potential to degrade antibiotics (Table 2).
Degradation of pesticides using white-rot fungi
In modern-day farming, pesticide application is more
common to increase crop produce and reduce post-harvest
losses (Hai et al. 2012). About 5% of applied pesticides
exterminate the specic pest organisms whereas the
remnants pass through surface and groundwater (Nawaz et
Tab le 2. Degradation of dierent antibiotics by various white-rot fungi
Antibiotics White-rot fungi Reference
Amoxicillin Trametes polyzona Lueangjaroenkit et al. 2019
Ampicillin Verticillium leptobactrum Kumar et al. 2013
Carbamazepine Trametes versicolor Hata et al. 2010a
Phanerochaete chrysosporium Zhang, Geißen 2010
Stropharia rugosoannulata, Gymnopilus luteofolius, Ganoderma lucidum,
Irpex lacteus, Agrocybe erebia
Castellet-Rovira et al. 2018
Chloramphenicol Trametes hirsuta Navada, Kulal 2019
Ciprooxacin Trametes versicolor Prieto et al. 2011
Pleurotus ostreatus Singh et al. 2017
Ganoderma lucidum Chakraborty, Abraham 2017
Pycnoporus sanguineus, Phanerochaete chrysosporium Gao et al. 2018
Xylaria longipes Rusch et al. 2018
Dichlofenac Phanerochaete sordida Hata et al. 2010b
Trametes trogii, Phanerochaete chrysosporium Aracagök et al. 2018
Pleurotus ostreatus Chapple et al. 2019
Erythromycin Trametes versicolor, Bjerkandera adusta Aydin et al. 2016
Ibuprofen Trametes versicolor, Irpex lacteus, Ganoderma lucidum, Phanerochaete
chrysosporium
Marco-Urrea et al. 2009
Lamotrigine Pleurotus ostreatus Chefetz et al. 2019
Naproxen Trametes versicolor Borràs et al. 2011
Noroxacin Trametes versicolor Prieto et al. 2011
Irpex lacteus, Panus tigrinus, Dichomitus squalens, Pleurotus ostreatus Čvančarová et al. 2015
Ganoderma lucidum Chakraborty, Abraham 2017
Ooxacin Trametes hirsute Haroune et al. 2014
Trametes versicolor, Irpex lacteus, Panus tigrinus, Dichomitus squalens,
Pleurotus ostreatus
Čvančarová et al. 2015
Sulfamethoxazole Phanerochaete chrysosporium Guo et al. 2014
Pleurotus ostreatus, Pleurotus pulmonarius, Trametes sp.de Araujo et al. 2017
Trametes versicolor Alharbi et al. 2019
Tetracycline Phanerochaete chrysosporium Wen et al. 2009
Trametes versicolor Suda et al. 2012
Cerrena sp. Yang et al. 2017
Pycnoporus sp. Tian et al. 2020
A. Kathiravan, J.J. Gnanadoss
111
al. 2011). e existence of agropesticides in the biosphere
has becomes a threat to ora, fauna, microbes and humans
(Hussain et al. 2015). e removal of lindane pesticides was
accomplished by Cyathus bulleri and Phanerochaete sordida.
Among these two species, Cyathus bulleri degraded more
eectively than Phanerochaete sordida. During the time
of degradation, two degradable intermediate metabolites
(tetrachlorocyclohexene and tetrachlorocyclohexanol)
were observed in Phanerochaete sordida culture.
Tetrachlorocyclohexanol was the rst detected degradation
product in Cyathus bulleri culture (Singh, Kuhad 2000).
Bending et al. (2002) studied the degradation potential
of nine white-rot fungi against monoaromatic pesticides
like diuron, metalaxyl, atrazine or terbuthylazine in liquid
culture. e highest level of pesticide degradation was
reported in Hypholoma fasciculare, Stereum hirsutum and
Coriolus ver sicolor. e rate of degradation of terbuthylazine,
diuron and atrazine was 86% while for metalaxyl the
degradation rate was below 44%. e capability of three
Phlebia species to degrade dieldrin and aldrin was also
examined (Xiao et al. 2011). Aer 42 days of treatment,
the three Phlebia sp. could degrade approximately 50% of
dieldrin in a low nitrogen medium. ree oxidized products
were identied as dieldrin metabolites in Phlebia species;
oxidation reactions might play an eective role in removing
dieldrin. Further, aldrin showed high degradation activity
and aer 28 days of culture, 90% of aldrin was degraded.
Transformed metabolites (two carboxylic acid products
and 9-hydroxyaldrin) were identied in the fungal cultures.
is showed that the methyl group of pesticides such as
aldrin and dieldrin might be susceptible to oxidative attack
by white-rot fungi. Clothianidin pesticide degradation
was tested using Phanerochaete sordida in nitrogen-
limited broth. Approximately 37% of clothianidin was
degraded at 30 °C aer an incubation period of 20 days.
N-(2-chlorothiazol-5-yl-methyl)-N’-methylurea was the
transformed metabolite during clothianidin degradation,
identied by analyzing the supernatant culture with high-
resolution electrospray ionisation mass spectrometry
and nuclear magnetic resonance (Mori et al. 2017). Some
common pesticides that are degraded by various white-rot
fungi are shown in Table 3.
Degradation of dyes using white-rot fungi
Articial dyes are broadly exploited in various industries
like food, cosmetics, pharmaceutical, textiles and leather,
etc. (Couto 2009). From 1856, over 105 dierent dyes
have been produced globally with a yearly production
of about 7 × 105 metric tons (Chen et al. 2003). Globally,
approximately 28 000 tons of textile dyestus are released
into textile industrial euent each year (Jin et al. 2007).
Unprocessed dye euents in water bodies cause severe
environmental and health threats (Shedbalkar et al. 2008).
Developing a cost-eective biological method to remove
synthetic colours is essential.
White-rot fungi are a class of fungi that synthesize
Tab le 3. Degradation of dierent pesticides by various white-rot fungi
Pesticide White-rot fungi Reference
Atrazine Pleurotus pulmonarius Masaphy et al. 1996
Anthracophyllum discolor Elgueta et al. 2016
Carbofuran Phlebia sp., Irpex lac teus Li et al. 2020
Chlorpyrifos Phlebia sp., Lenzites betulinus, Irpex lacteus Wang et al. 2020
Clothianidin Phanerochaete sordida Mori et al. 2017
Dichlorophen Bjerkandera adusta Davila-Vazquez et al. 2005
Dichlorophenoyacetcid Lentinula edodes Tsujiyama et al. 2013
Lentinus crinitus Serbent et al. 2020
Diuron Agrocybe semiorbicularis, Auricularia auricola, Flammulina
velupites, Dichotomitus squalens, Coriolus veriscolor, Hypholoma
fasciculare, Phanerochaete velutina, Pleurotus ostreatus, Stereum
hirsutum
Bending et al. 2002
Ceriporia lacerata, Phanerochaete chrysosporium,
Phanerochaete sordida, Trametes versicolor
Mori et al. 2018
Endrin Phlebia acanthocystis, Phlebia brevispora Xiao, Kondo 2019
Fipronil Trametes versicolor Wolfand et al. 2016
Lindane Cyathus bulleri, Phanerochaete sordida Singh, Kuhad 2000
Ganoderma australe Dritsa et al. 2005
Pleurotus ostreatus Papadopoulou et al. 2006
Ganoderma lucidum Kaur et al. 2016
Parathion Bjerkandera adusta, Pleurotus ostreatus, Phanerochaete
chrysosporium
Jauregui et al. 2003
1,1’-(2,2,2-Trichloroethane-1,1-
diyl)bis(4-chlorobenzene) (DDT)
Phlebia lindineri, Phlebia brevispora Xiao et al. 2011
White-rot fungi mediated bioremediation for xenobiotic degradation
112
enzymes able to decompose dyes in aerobic conditions
(Nozaki et al. 2008). ey produce several oxidoreductases
that can biodegrade lignin and their associated aromatic
compounds. e capacity for dye degradation diers for
fungal species and enzymes (Nyanhongo et al. 2002). Four
dierent mechanisms are involved in the decolouration of
dye using white-rot fungi: biodegradation, biosorption,
bioreactor and immobilized lignin modied enzymes
(Jebapriya, Gnanadoss 2013). e biosorption mechanism
involves the adsorption of dyes by fungal biomass. However,
dye removal by adsorption was found to be restricted to
50% (Knapp et al. 2001). Biodegradation has a key role in
dye decolourization, as it secretes extracellular lignolytic
enzymes to oxidize the dyes (Jayasinghe et al. 2008).
Laccase producing white-rot fungi Pleurotus floridanus
LCJ155, Leucocoprinus cretaceous LCJ164, and Agaricus sp.
LCJ169 were eective in degrading synthetic dyes such as
bromophenol blue, methyl red, phenol red, Congo red and
brilliant green (Jebapriya, Gnanadoss 2014). Bjerkandera
adusta cultured in potato dextrose broth medium in an
airli bioreactor. Aer 10- to 15-h treatment, there was 90%
of dye decolourization for both acid and reactive colourants.
ese results suggested that a bioreactor employed with a
white-rot fungal strain is promising for dye euent removal
(Sodaneath et al. 2017). e decolourization of erichrome
black T and Congo red dyes by Pleurotus ostreatus was
studied (Gnanadoss et al. 2013). e highest degradation
rate of dyes was observed when Pleurotus ostreatus culture
was immobilized on polyurethane foam.
e LiP enzyme obtained from Ganoderma lucidum
(GRM117) and Pleurotus ostreatus (PLO9) immobilized
on carbon nanotubes is a promising biocatalyst for dye
removal (Oliveira et al. 2018). Biosorption of remazol
brilliant blue R and indigo carmine dyes was studied using
immobilized biomass of white-rot fungus Psathyrella
candolleana LCJ178. Polyurethane foam, stainless steel
sponge, lua sponge, scotch brite and white nylon sponge
were used as supporting materials for immobilization
of Psathyrella candolleana LCJ178 biomass. Of these,
stainless steel sponge was found eective in binding to
the culture without causing any operational diculties.
is study revealed that the selection of suitable support
material, culture condition (shaking) and other physical
aspects were critical for enhancing the process of dye
removal (Gnanasalomi et al. 2016). Several reviews on
the dye removal by means of white-rot fungi have already
been published (Shah, Nerud 2002; Wesenberg et al. 2003;
Asgher et al. 2008; Tišma et al. 2010; Jebapriya, Gnanadoss
2013; Sen et al. 2016; Chaturvedi et al. 2019; Periasamy et
al. 2019).
Conclusions
Xenobiotic compounds are anthropogenic substances
generated from multiple industries that have negative
environmental consequences if they are released without
proper pretreatment. Xenobiotics are noxious to living
organisms; therefore they need to be removed quickly
before entering into the environment. However, the
physical and chemical methodologies are not feasible
enough to degrade xenobiotic compounds. Subsequently,
an alternative remediation technology is needed to combat
xenobiotic compounds. Bioremediation technology is
more promising in xenobiotic degradation owing to its
cost-eective and eco-friendly approach.
White-rot fungi are thought to be ecient bio-degraders
of organic pollutants probably owing to their metabolic
enzymes with extensive substrate specicities. Dierent
white-rot fungi have dierent biodegradation abilities for
dierent xenobiotic compounds primarily due to their
unique morphology, culture and environmental aspects as
well as the nature of the enzymes produced. e characteristic
features of lignolytic enzymes dier between taxa of white-
rot fungi. ey are well explored in the biodegradation of
diverse xenobiotics like dyes, hydrocarbons, and phenolic
compounds on a laboratory scale. Still, many studies are
required to explore the scope of white-rot fungi at the
industrial level. Additionally, screening of new white-rot
fungal isolates oen with promising enzyme activity is
required for the bioremediation of new toxic chemicals due
to increased industrial contamination. White-rot fungi in
combination with biotechnological tools such as genetic
engineering can produce novel strains with ideal properties
for the disintegration of numerous xenobiotic pollutants.
Acknowledgements
e authors are grateful to the management of Loyola College,
Chennai for providing necessary facilities and encouragement.
References
Abdel-Shafy H.I., Mansour M.S. 2016. A review on polycyclic
aromatic hydrocarbons: source, environmental impact, eect
on human health and remediation. Egypt. J. Petrol. 25: 107–
123.
Acevedo F., Pizzul L., del Pilar Castillo M., Cuevas R., Diez M.C.
2011. Degradation of polycyclic aromatic hydrocarbons by
the Chilean white-rot fungus Anthracophyllum discolor. J.
Hazard. Mater. 185: 212–219.
Adenipekun C.O., Fasidi I.O. 2005. Bioremediation of oil-polluted
soil by Lentinus subnudus, a Nigerian white-rot fungus. Afr. J.
Biotechnol. 4: 796–798.
Agrawal N., Shahi S.K. 2017. Degradation of polycyclic aromatic
hydrocarbon (pyrene) using novel fungal strain Coriolopsis
byrsina strain APC5. Int. Biodeterior. Biodegrad. 122: 69–81.
Agrawal N., Verma P., Singh R.S., Shahi S.K. 2017. Ligninolytic
enzyme production by white rot fungi Podoscypha elegans
strain FTG4. Int. J. Curr. Microbiol. Appl. Sci. 6: 2757–2764.
Agrawal N., Verma P., Shahi S.K. 2018. Degradation of polycyclic
aromatic hydrocarbons (phenanthrene and pyrene) by the
ligninolytic fungi Ganoderma lucidum isolated from the
hardwood stump. Bioresour. Bioprocess. 5: 1–9.
A. Kathiravan, J.J. Gnanadoss
113
Aksu Z. 2005. Application of biosorption for the removal of
organic pollutants: a review. Process. Biochem. 40: 997–1026.
Alharbi S.K., Nghiem L.D., Van De Merwe J.P., Leusch F.D., Asif
M.B., Hai F.I., Price W.E. 2019. Degradation of diclofenac,
trimethoprim, carbamazepine, and sulfamethoxazole by
laccase from Trametes versicolor: Transformation products
and toxicity of treated euent. Biocatal. Biotransform. 37:
399–408.
Amara S., Perrot T., Navarro D., Deroy A., Benkhelfallah A.,
Chalak A., Morel-Rouhier M. 2018. Enzyme activities of two
recombinant heme-containing peroxidases, TvDyP1 and
TvVP2, identied from the secretome of Trametes versicolor.
Appl. Environ. Microbiol. 84: e02826-17.
Andlar M., Rezić T., Marđetko N., Kracher D., Ludwig R., Šantek
B. 2018. Lignocellulose degradation: an overview of fungi and
fungal enzymes involved in lignocellulose degradation. Eng.
Life Sci. 18: 768–778.
Aracagök Y.D., Göker H., Cihangir N. 2018. Biodegradation of
diclofenac with fungal strains. Arch. Environ. Prot. 44: 55–62.
Aranda E., Kinne M., Kluge M., Ullrich R., Hofrichter M. 2009.
Conversion of dibenzothiophene by the mushrooms Agrocybe
aegerita and Coprinellus radians and their extracellular
peroxygenases. Appl. Microbiol. Biotechnol. 82: 1057–1066.
Arora D.S., Gill P.K. 2001. Eects of various media and
supplements on laccase production by some white rot fungi.
Bioresour. Technol. 77: 89–91.
Asgher M., Bhatti H.N., Ashraf M., Legge R.L. 2008. Recent
developments in biodegradation of industrial pollutants by
white-rot fungi and their enzyme system. Biodegradation 19:
771–783.
Aydin S. 2016. Enhanced biodegradation of antibiotic
combinations via the sequential treatment of the sludge
resulting from pharmaceutical wastewater treatment using
white-rot fungi Trametes versicolor and Bjerkandera adusta.
Appl. Microbiol. Biotechnol. 100: 6491–6499.
Baborová P., Möder M., Baldrian P., Cajthamlová K., Cajthaml
T. 2006. Purication of a new manganese peroxidase of the
white-rot fungus Irpex lacteus and degradation of polycyclic
aromatic hydrocarbons by the enzyme. Res. Microbiol. 157:
248–253.
Baker P., Tiroumalechetty A., Mohan R. 2019. Fungal enzymes for
bioremediation of xenobiotic compounds. In: Yadav A.N. et
al. (Eds.) Recent Advancement in White Biotechnology rough
Fungi. Springer Nature Switzerland, pp. 463–489.
Bautista L.F., Morales G., Sanz, R. 2010. Immobilization strategies
for laccase from Trametes versicolor on mesostructured
silica materials and the application to the degradation of
naphthalene. Bioresour. Technol. 101: 8541–8548.
Bending G.D., Friloux M., Walker A. 2002. Degradation of
contrasting pesticides by White-rot fungi and its relationship
with ligninolytic potential. FEMS Microbiol. Lett. 212: 59–63.
Bishnoi K., Kumar R., Bishnoi N.R. 2008. Biodegradation
of polycyclic aromatic hydrocarbons by white-rot fungi
Phanerochaete chrysosporium in sterile and unsterile soil. J.
Sci . Ind. Res. 67: 538–542.
Borja J., Taleon D.M., Auresenia J., Gallardo S. 2005.
Polychlorinated biphenyls and their biodegradation. Process
Biochem. 40: 1999–2013.
Borràs E., Llorens-Blanch G., Rodríguez-Rodríguez C.E., Sarrà
M., Caminal G. 2011. Soil colonization by Trametes versicolor
grown on lignocellulosic materials: substrate selection and
naproxen degradation. Int. Biodeterior. Biodegrad. 65: 846–
852.
Bouki, C., Venieri, D., Diamadopoulos E. 2013. Detection and fate
of antibiotic resistant bacteria in wastewater treatment plants:
a review. Ecotoxicol. Environ. Safety 91: 1–9.
Bressler D.C., Fedorak P.M., Pickard M.A. 2000. Oxidation of
carbazole, N-ethylcarbazole, uorene, and dibenzothiophene
by the laccase of Coriolopsis gallica. Biotechnol. Lett. 22: 1119–
1125.
Bumpus J.A., Aust S.D. 1987. Biodegradation of environmental
pollutants by the white rot fungus Phanerochaete
chrysosporium: involvement of the lignin degrading system.
Bioessays 6: 166–170.
Castellet-Rovira F., Lucas D., Villagrasa M., Rodríguez-Mozaz
S., Barceló D., Sarrà M. 2018. Stropharia rugosoannulata
and Gymnopilus luteofolius: Promising fungal species for
pharmaceutical biodegradation in contaminated water. J.
Environ. Manage. 207: 396–404.
Chakraborty P., Abraham J. 2017. Comparative study on
degradation of noroxacin and ciprooxacin by Ganoderma
lucidum JAPC1. Korean J. Chem. Eng. 34: 1122–1128.
Chaplin A.K., Chicano T.M., Hampshire B.V., Wilson M.T., Hough
M.A., Svistunenko D.A., Worrall J.A. 2019. An aromatic dyad
motif in dye decolourising peroxidases has implications for
free radical formation and catalysis. Chem. Eur. J. 25: 6141–
6153.
Chapple A., Nguyen L.N., Hai F.I., Dosseto A., Rashid M.H.O., Oh
S., Nghiem L.D. 2019. Impact of inorganic salts on degradation
of bisphenol A and diclofenac by crude extracellular enzyme
from Pleurotus ostreatus. Biocatal. Biotransf. 37: 10–17.
Chaturvedi S. 2019. Azo dyes decolorization using white-rot
fungi. Res. Rev. J. Microbiol. Biotechnol. 8: 9–19.
Chefetz B., Marom R., Salton O., Oliferovsky M., Mordehay V.,
Ben-Ari J., Hadar Y. 2019. Transformation of lamotrigine
by white-rot fungus Pleurotus ostreatus. Environ. Pollut. 250:
546–553.
Chen K.C., Wu J.Y., Liou D.J., Hwang S.C.J. 2003. Decolorization
of the textile dyes by newly isolated bacterial strains. J.
Biotechnol. 101: 57–68.
Chen J., Hu J. D., Wang X. 2005. Degradation of polycyclic
aromatic hydrocarbons from soil by white-rot fungi. Environ.
Chem. 24: 274.
Cheong S., Yeo S., Song H.G., Choi H.T. 2006. Determination
of laccase gene expression during degradation of
2,4,6-trinitrotoluene and its catabolic intermediates in
Trametes versicolor. Microbiol. Res. 161: 316–320.
Christian V., Shrivastava R., Shukla D., Modi H.A., Vyas B.R.M.
2005. Degradation of xenobiotic compounds by lignin-
degrading white-rot fungi: enzymology and mechanisms
involved. Indian J. Exp. Biol. 43: 301.
Chupungars K., Rerngsamran P., aniyavarn S. 2009. Polycyclic
aromatic hydrocarbons degradation by Agrocybe sp. CU-43
and its uorene transformation. Int. Biodeterior. Biodegrad.
63: 93–99.
Colvin S., Nelson, D. 1990. White-rot fungi in PCB biodegradation.
In: Proceedings of hazardous waste management in the 90’s:
moving from remediation to practical preventive strategies.
A&WMA, Pittsburgh.
Couto S.R. 2009. Dye removal by immobilised fungi. Biotechnol.
Adv. 27: 227–235.
Čvančarová M., Moeder M., Filipová A., Cajthaml T. 2015.
White-rot fungi mediated bioremediation for xenobiotic degradation
114
Biotransformation of uoroquinolone antibiotics by
ligninolytic fungi: Metabolites, enzymes and residual
antibacterial activity. Chemosphere 136: 311–320.
Cycoń M., Mrozik A., Piotrowska-Seget Z. 2019. Antibiotics in the
soil environment-degradation and their impact on microbial
activity and diversity. Front. Microbiol. 10: 338.
Czaplicka M. 2004. Sources and transformations of chlorophenols
in the natural environment. Sci. Total Environ. 322: 21–39.
da Silva Coelho-Moreira J., Brugnari T., Sá-Nakanishi A.B.,
Castoldi R., de Souza C.G., Bracht A., Peralta R.M. 2018.
Evaluation of diuron tolerance and biotransformation by
the white-rot fungus Ganoderma lucidum. Fungal Biol. 122:
471–478.
Das M.P., Kumar S. 2014. Microbial deterioration of low density
polyethylene by Aspergillus and Fusarium sp. Int. J. Chem.
Tech. R es . 6: 299–305.
Dashtban M., Schra H., Syed T.A., Qin W. 2010. Fungal
biodegradation and enzymatic modication of lignin. Int. J.
Biochem. Mol. Biol. 1: 36–50.
Davila-Vazquez G., Tinoco R., Pickard M.A., Vazquez-Duhalt R.
2005. Transformation of halogenated pesticides by versatile
peroxidase from Bjerkandera adusta. Enz. Microb. Technol. 36:
223–231.
de Araujo C.A.V., Maciel G.M., Rodrigues E.A., Silva L.L., Oliveira
R.F., Brugnari T., de Souza C.G.M. 2017. Simultaneous removal
of the antimicrobial activity and toxicity of sulfamethoxazole
and trimethoprim by white-rot fungi. Water Air Soil Pollut.
228: 341.
Deshmukh R., Khardenavis A.A., Purohit H.J. 2016. Diverse
metabolic capacities of fungi for bioremediation. Indian J.
Microbiol. 56: 247–264.
Dhiman N., Jasrotia T., Sharma P., Negi S., Chaudhary S., Kumar
R., Kumar R. 2020. Immobilization interaction between
xenobiotic and Bjerkandera adusta for the biodegradation of
atrazine. Chemosphere 257: 127060.
Ding C., Wang X., Li M. 2019. Evaluation of six white-rot fungal
pretreatments on corn stover for the production of cellulolytic
and ligninolytic enzymes, reducing sugars, and ethanol. Appl.
Microbiol. Biotechnol. 103: 5641–5652.
Donnelly K.C., Chen J.C., Huebner H.J., Brown K.W., Autenrieth
R.L., Bonner J.S. 1997. Utility of four strains of white‐rot
fungi for the detoxication of 2, 4, 6‐trinitrotoluene in liquid
culture. Environ. Toxicol. Chem. 16: 1105–1110.
Dosoretz C.G., Reddy C.A. 2007. Lignin and lignin-modifying
enzymes. In: Methods for General and Molecular Microbiology.
3rd. Ed. American Society of Microbiology. pp. 611–620.
Dritsa V., Rigas F., Avramides E.J., Hatzianestis I. 2005.
Biodegradation of lindane in liquid cultures by the
polypore fungus Ganoderma australe. In: ird European
Bioremediation Conference, Chania, Greece.
Dwivedi U.N., Singh P., Pandey V.P., Kumar A. 2011. Structure–
function relationship among bacterial, fungal and plant
laccases. J. Mol. Catal. B Enzym. 68: 117–128.
Eaton D.C. 1985. Mineralization of polychlorinated biphenyls by
Phanerochaete chrysosporium: a ligninolytic fungus. Enzyme
Microb. Technol. 7: 194–196.
Eaton R.A., Hale M.D. 1993. Wood: Decay, Pests and Protection.
Chapman and Hall.
Elgueta S., Santos C., Lima N., Diez M.C. 2016. Immobilization of
the white-rot fungus Anthracophyllum discolor to degrade the
herbicide atrazine. AMB Express 6: 104.
Ellouze M., Sayadi S. 2016. White-rot fungi and their enzymes as a
biotechnological tool for xenobiotic bioremediation. In: Saleh
H., Rahman R.A. (Eds.) Management of Hazardous Wastes.
InTech, London, pp.103–120.
Erden E., Ucar M.C., Gezer T., Pazarlioglu N.K. 2009. Screening
for ligninolytic enzymes from autochthonous fungi and
applications for decolorization of Remazole Marine Blue.
Braz. J. Microbiol. 40: 346–353.
Faraco V., Piscitelli A., Sannia G., Giardina P. 2007. Identication
of a new member of the dye-decolorizing peroxidase family
from Pleurotus ostreatus. World. J. Microb. Biotechnol. 23:
889–893.
Fernández-Fueyo E., Linde D., Almendral D., López-Lucendo
M.F., Ruiz-Dueñas F.J., Martínez A.T. 2015. Description of the
rst fungal dye-decolorizing peroxidase oxidizing manganese
(II). Appl. Microbiol. Biotechnol. 99: 8927–8942.
Field J.A., De Jong E., Costa G.F., De Bont J.A. 1992. Biodegradation
of polycyclic aromatic hydrocarbons by new isolates of white-
rot fungi. Appl. Environ. Microbiol. 58: 2219–2226.
Field J.A., de Jong E., Feijoo-Costa G., de Bont J.A. 1993. Screening
for ligninolytic fungi applicable to the biodegradation of
xenobiotics. Trends Biotechnol. 11: 44–49.
Field J.A., Sierra-Alvarez R. 2008. Microbial degradation of
chlorinated phenols. Rev. Environ. Sci. Biotechnol. 7: 211–241.
Friedrich J., Zalar P., Mohorčič M., Klun U., Kržan A. 2007. Ability
of fungi to degrade synthetic polymer nylon-6. Chemosphere
67: 2089–2095.
Gao N., Liu C.X., Xu Q.M., Cheng J.S., Yuan Y.J. 2018. Simultaneous
removal of ciprooxacin, noroxacin, sulfamethoxazole by
co-producing oxidative enzymes system of Phanerochaete
chrysosporium and Pycnoporus sanguineus. Chemosphere 195:
146–155.
Gelband H., Molly Miller P., Pant S., Gandra S., Levinson J., Barter
D., Laxminarayan R. 2015. e state of the world’s antibiotics
2015. Wound Healing Southern Africa 8: 30–34.
Gianfreda L., Rao M.A. 2008. Interactions between xenobiotics
and microbial and enzymatic soil activity. Crit. Rev. Environ.
Sci. Technol. 38: 269–310.
Gill P.K.D.A., Arora D. 2003. Eect of culture conditions on
manganese peroxidase production and activity by some white
rot fungi. J. Industr. Microbiol. Biotechnol. 30: 28–33.
Gnanadoss J.J., Peeris S.M., Jebapriya G.R. 2013. Biologically
enhanced absorption of synthetic dyes using oyster mushroom
fungus Pleurotus ostreatus. Asian J. Exp. Biol. Sci. 4: 637–640.
Gnanasalomi V.D.V., Gnanadoss J.J. 2013. Laccases from fungi
and their applications: recent developments. Asian J. Exp. Biol.
Sci . 4: 581–590.
Gnanasalomi V.D.V., Gnanadoss J.J. 2013. Molecular
characterization and phylogenetic analysis of laccase
producing fungal isolates with dye decolourizing potential.
Res. Biotechnol. 4: 1–8.
Gnanasalomi V.D.V., Gnanadoss J.J. 2019. Laccase production by
Myrothecium gramineum and its optimization under solid
state fermentation using cowpea pod as substrate. J. Appl . Biol.
Biotechnol. 7: 59–63.
Gnanasalomi V.D.V., Jebapriya G.R., Gnanadoss J.J. 2013.
Bioremediation of hazardous pollutants using fungi. Int. J.
Comput. Algor. 2: 93–96.
Gnanasalomi V.D.V., Priya E.S., Gnanadoss J.J. 2016. Biosorption
of remazol brilliant blue R and indigo carmine using
immobilized biomass of Psathyrella candolleana LCJ178.
A. Kathiravan, J.J. Gnanadoss
115
Pollut. Res. 35: 337–342.
Gren I. 2012. Microbial transformation of xenobiotics. Chemik 66:
839–842.
Guo X.L., Zhu Z.W., Li H.L. 2014. Biodegradation of
sulfamethoxazole by Phanerochaete chrysosporium. J. Mol.
Liquids 198: 169–172.
Gursahani Y.H., Gupta S.G. 2011. Decolourization of textile
euent by a thermophilic bacteria Anoxybacillus rupiensis. J.
Pet. Environ. Biotechnol. 2: 1–4.
Hadibarata T., Tachibana S., Itoh K. 2009. Biodegradation of
chrysene, an aromatic hydrocarbon by Polyporus sp. S133 in
liquid medium. J. Hazard. Mater. 164: 911–917.
Hadibarata T., Yuso A.R.M., Aris A., Kristanti R.A. 2012.
Identication of naphthalene metabolism by white rot fungus
Armillaria sp. F022. J. Environ. Sci. 24: 728–732.
Hadibarata T., Kristanti R.A. 2014. Potential of a white-rot fungus
Pleurotus eryngii F032 for degradation and transformation of
uorene. Fungal. Biol. 118: 222–227.
Hai F.I., Modin O., Yamamoto K., Fukushi K., Nakajima F.,
Nghiem L.D. 2012. Pesticide removal by a mixed culture of
bacteria and white-rot fungi. J. Taiwan. Inst. Chem. Eng. 43:
459–462.
Hakala T.K., Lundell T., Galkin S., Maijala P., Kalkkinen N.,
Hatakka A. 2005. Manganese peroxidases, laccases and
oxalic acid from the selective white-rot fungus Physisporinus
rivulosus grown on spruce wood chips. Enzyme Microb.
Techn ol. 36: 461–468.
Hammel K.E., Cullen D. 2008. Role of fungal peroxidases in
biological ligninolysis. Curr. Opin. Plant Biol. 11: 349–355.
Han M.J., Choi H.T., Song H.G. 2004. Degradation of phenanthrene
by Trametes versicolor and its laccase. J. Microbiol. 42: 94–98.
Haroune L., Saibi S., Bellenger J.P., Cabana H. 2014. Evaluation of
the eciency of Trametes hirsuta for the removal of multiple
pharmaceutical compounds under low concentrations
relevant to the environment. Bioresour. Technol. 171: 199–202.
Hata T., Shintate H., Kawai S., Okamura H., Nishida T. 2010a.
Elimination of carbamazepine by repeated treatment with
laccase in the presence of 1-hydroxybenzotriazole. J. Hazard.
Mate r. 181: 1175–1178.
Hata T., Kawai S., Okamura H., Nishida T. 2010b. Removal of
diclofenac and mefenamic acid by the white rot fungus
Phanerochaete sordida YK-624 and identication of their
metabolites aer fungal transformation. Biodegradation 21:
681–689.
Heberer T. 2002. Occurrence, fate, and removal of pharmaceutical
residues in the aquatic environment: a review of recent
research data. Toxicol. Lett. 131: 5–17.
Hofrichter M. 2002. Lignin conversion by manganese peroxidase
(MnP). Enzyme Microb. Technol. 30: 454–466.
Honeycutt M.E., Jarvis A.S., McFarland V.A. 1996. Cytotoxicity
and mutagenicity of 2,4,6-trinitrotoluene and its metabolites.
Ecotoxicol. Environ. Safety 35: 282–287.
Hong C.Y., Gwak K.S., Lee S.Y., Kim S.H., Lee S.M., Kwon M.,
Choi I.G. 2012. Biodegradation of PCB congeners by white rot
fungus, Ceriporia sp. ZLY-2010, and analysis of metabolites. J.
Environ. Sci. Health A 47: 1878–1888.
Hussain S., Arshad M., Springael D., Sørensen S.R., Bending G.D.,
Devers-Lamrani M., Martin-Laurent F. 2015. Abiotic and
biotic processes governing the fate of phenylurea herbicides in
soils: a review. Crit. Rev. Env iron. Sci. Technol. 45: 1947–1998.
Ichinose H. 2013. Cytochrome P 450 of wood‐rotting
basidiomycetes and biotechnological applications. Biotechnol.
Appl. Biochem. 60: 71–81.
Jauregui J., Valderrama B., Albores A., Vazquez-Duhalt R. 2003.
Microsomal transformation of organophosphorus pesticides
by white-rot fungi. Biodegradation 14: 397–406.
Jayasinghe C., Imtiaj A., Lee G.W., Im K.H., Hur H., Lee M.W.,
Lee T.S. 2008. Degradation of three aromatic dyes by white-
rot fungi and the production of ligninolytic enzymes.
Mycobiology 36: 114–120.
Jebapriya G.R., Gnanadoss J.J. 2013. Bioremediation of textile dye
using white-rot fungi: A review. Int. J. Curr. Res. 5: 1.
Jebapriya G.R., Gnanadoss J.J. 2014. Screening and molecular
characterization of white rot fungi capable of laccase
production and dye decolourization. Int. J. Life Sci. Pharma
Res. 4: 12–20.
Jin X.C., Liu G.Q., Xu Z.H., Tao W.Y. 2007. Decolorization of a
dye industry euent by Aspergillus fumigatus XC6. Appl.
Microbiol. Biotechnol. 74: 239–243.
Johjima T., Ohkuma M., Kudo T. 2003. Isolation and cDNA
cloning of novel hydrogen peroxide-dependent phenol
oxidase from the basidiomycete Termitomyces albuminosus.
Appl. Microbiol. Biotechnol. 61: 220–225.
Jose L.J., Joel G.J. 2014. Optimization of culture conditions for
improved laccase production by Agaricu s sp. LCJ262. Int. J.
Curr. R es. 6: 9517–9522.
Kale S.K., Deshmukh A.G., Dudhare M.S., Patil V.B. 2015.
Microbial degradation of plastic: a review. J. Biochem. Technol.
6: 952–961.
Kaseem M., Hamad K., Deri F. 2012. ermoplastic starch blends:
A review of recentworks. Polym. Sci. Ser. A 54: 165–176.
Kaur H., Kapoor S., Kaur G. 2016. Application of ligninolytic
potentials of a white-rot fungus Ganoderma lucidum for
degradation of lindane. Environ. Monit. Assess. 188: 588.
Kılıç N., Nasiri F., Cansaran-Duman D. 2016. Fungal laccase
enzyme applications in bioremediation of polluted
wastewater. In: Ansari A.A. et al. (Eds.) Phytoremediation.
Springer Nature, pp. 201–209.
Kim H.Y., Song H.G. 2000. Comparison of 2,4,6-trinitrotoluene
degradation by seven strains of white-rot fungi. Curr.
Microbiol. 41: 317–320.
Kim H.Y., Song H.G. 2003. Transformation and mineralization
of 2,4,6-trinitrotoluene by the white rot fungus Irpex lacteus.
Appl. Microbiol. Biotechnol. 61: 150–156.
Knapp J.S., Newby P.S. 1995. e microbiological decolorization of
an industrial euent containing a diazo-linked chromophore.
Water. Res. 29: 1807–1809.
Kolwek J., Behrens C., Linke D., Krings U., Berger R.G. 2018. Cell-
free one-pot conversion of (+)-valencene to (+)-nootkatone
by a unique dye-decolorizing peroxidase combined with a
laccase from Funalia trogii. J. Indian Microbiol. Biotechnol. 45:
89–101.
Krahe N.K., Berger R.G., Ersoy F. 2020. A DyP-Type peroxidase
of Pleurotus sapidus with alkene cleaving activity. Molecules
25: 1536.
Kumar R.R., Park B.J., Jeong H.R., Lee J.T., Cho J.Y. 2013.
Biodegradation of β-lactam antibiotic ‘Ampicillin’ by white-
rot fungi from aqueous solutions. J. Pure Appl. Microbiol. 7:
3163–3169.
Kunjadia P.D., Sanghvi G.V., Kunjadia A.P., Mukhopadhyay P.N.,
Dave G.S. 2016. Role of ligninolytic enzymes of white rot
fungi (Pleurotus spp.) grown with azo dyes. SpringerPlus 5:
White-rot fungi mediated bioremediation for xenobiotic degradation
116
1487.
Lazim Z.M., Hadibarata T. 2016. Ligninolytic fungus Polyporus
sp. S133 mediated metabolic degradation of uorene. Braz. J.
Microbiol. 47: 610–616.
Lee A. H., Lee H., Kim J.J. 2016. Simultaneous degradation of
polycyclic aromatic hydrocarbons by attractive ligninolytic
enzymes from Phlebia brevispora KUC9045. Korean J. Enviro n.
Biol. 34: 201–207.
Levin L., Papinutti L., Forchiassin F. 2004. Evaluation of
Argentinean white rot fungi for their ability to produce lignin-
modifying enzymes and decolorize industrial dyes. Bioresour.
Techn ol . 94: 169–176.
Li N., Xia Q., Niu M., Ping Q., Xiao H. 2018. Immobilizing laccase
on dierent species wood biochar to remove the chlorinated
biphenyl in wastewater. Sci. Rep. 8: 13947.
Li X., Pan Y., Hu S., Cheng Y., Wang Y., Wu K., Yang S. 2018.
Diversity of phenanthrene and benz[a]anthracene metabolic
pathways in white rot fungus Pycnoporus sanguineus 14. Int.
Biodeterior. Biodegrad. 134: 25–30.
Li Z., Wang X., Ni Z., Bao J., Zhang H. 2020. In-situ remediation
of carbofuran-contaminated soil by immobilized white-rot
fungi. Pol. J. Environ. Stud. 29: 1237–1243.
Liers C., Pecyna M.J., Kellner H., Worrich A., Zorn H., Steen
K.T., Ullrich R. 2013. Substrate oxidation by dye-decolorizing
peroxidases (DyPs) from wood-and litter-degrading
agaricomycetes compared to other fungal and plant heme-
peroxidases. Appl. Microbiol. Biotechnol. 97: 5839–5849.
Lueangjaroenkit P., Teerapatsakul C., Sakka K., Sakka M.,
Kimura T., Kunitake E., Chitradon L. 2019. Two manganese
peroxidases and a laccase of Trametes polyzona KU-RNW027
with novel properties for dye and pharmaceutical product
degradation in redox mediator-free system. Mycobiology. 47:
217–229.
Marco-Urrea E., Pérez-Trujillo M., Vicent T., Caminal G. 2009.
Ability of white-rot fungi to remove selected pharmaceuticals
and identication of degradation products of ibuprofen by
Trametes versicolor. Chemosphere 74: 765–772.
Martens R., Zadrazil F. 1996. Two-step degradation of pyrene by
white-rot fungi and soil microorganisms. Appl. Microbiol.
Biotechnol. 46: 653–659.
Masaphy S., Henis Y., Levanon D. 1996. Manganese-enhanced
biotransformation of atrazine by the white rot fungus
Pleurotus pulmonarius and its correlation with oxidation
activity. Appl. Environ. Microbiol. 62: 3587–3593.
Mester T., Field J.A. 1997. Optimization of manganese peroxidase
production by the white rot fungus Bjerkandera sp. strain
BOS55. FEMS Microbiol. Lett. 155: 161–168.
Mishra V.K., Singh G., Shukla R. 2019. Impact of xenobiotics
under a changing climate scenario. In: Choudhary K.K.,
Kumar A., Singh A.K. (Eds.) Climate Change and Agricultural
Ecosystems: Current Challenges and Adaptation. Woodhead
Publishing, pp. 133–151
Mohammadi-Sichani M., Assadi M.M., Farazmand A., Kianirad
M., Ahadi A.M., Hadian-Ghahderijani H. 2019. Ability of
Agaricus bisporus, Pleurotus ostreatus and Ganoderma lucidum
compost in biodegradation of petroleum hydrocarbon-
contaminated soil. Int. J. Environ. Sci. Technol. 16: 2313–2320.
Mori T., Kitano S., Kondo R. 2003. Biodegradation of
chloronaphthalenes and polycyclic aromatic hydrocarbons
by the white-rot fungus Phlebia lindtneri. Appl. Microbiol.
Biotechnol. 61: 380–383.
Mori T., Wang J., Tanaka Y., Nagai K., Kawagishi H., Hirai H. 2017.
Bioremediation of the neonicotinoid insecticide clothianidin
by the white-rot fungus Phanerochaete sordida. J. Hazard.
Mate r. 321: 586–590.
Mori T., Sudo S., Kawagishi H., Hirai H. 2018. Biodegradation
of diuron in articially contaminated water and seawater by
wood colonized with the white-rot fungus Trametes versicolor.
J. Wood Sci. 64: 690–696.
Navada K., Kulal A. 2019. Enzymatic degradation of
chloramphenicol by laccase from Trametes hirsuta and
comparison among mediators. Int. Biodeterior. Biodegrad.
138: 63–69.
Nawaz K., Hussain K., Choudary N., Majeed A., Ilyas U., Ghani
A., Lashari M.I. 2011. Eco-friendly role of biodegradation
against agricultural pesticides hazards. Afr. J. Microbiol. Res.
5: 177–183.
Nerud F., Mišurcová Z. 1996. Distribution of ligninolytic enzymes
in selected white-rot fungi. Folia Microbiol. 41: 264–266.
Nikiforova S.V., Pozdnyakova N.N., Makarov O.E., Chernyshova
M.P., Turkovskaya O.V. 2010. Chrysene bioconversion by
the white rot fungus Pleurotus ostreatus D1. Microbiology 79:
456–460.
Nozaki K., Beh C.H., Mizuno M., Isobe T., Shiroishi M., Kanda
T., Amano Y. 2008. Screening and investigation of dye
decolorization activities of basidiomycetes. J. Biosci. Bioeng.
105: 69–72.
Oliveira S. F., da Luz J.M.R., Kasuya M.C.M., Ladeira L.O., Junior
A.C. 2018. Enzymatic extract containing lignin peroxidase
immobilized on carbon nanotubes: Potential biocatalyst in
dye decolourization. Saudi J. Biol. S ci. 25: 651–659.
Papadopoulou K., Rigas F., Doulia D. 2006. Lindane degradation
in soil by Pleurotus ostreatus. WSEAS Trans. Environ. Dev. 2:
489–496.
Park H., Min, B., Jang Y., Kim J., Lipzen A., Sharma A., Henrissat
B. 2019. Comprehensive genomic and transcriptomic
analysis of polycyclic aromatic hydrocarbon degradation by
a mycoremediation fungus, Dentipellis sp. KUC8613. Appl.
Microbiol. Biotechnol. 103: 8145–8155.
Paszczyński A., Huynh V.B., Crawford R. 1985. Enzymatic
activities of an extracellular, manganese-dependent peroxidase
from Phanerochaete chrysosporium. FEMS Microbiol. Lett. 29:
37–41.
Periasamy D., Mani S., Ambikapathi R. 2019. White-rot fungi and
their enzymes for the treatment of industrial dye euents.
In: Yadav A.N. et al. (Eds.) Recent Advancement in White
Biotechnology rough Fungi. Springer Nature Switzerland,
pp. 73–100.
Perkins, M.W., De S., Frederick L., Dutta S.K. 2005. Ligninolytic
and nonligninolytic mineralization of trinitrotoluene by
several white rot basidiomycetes. Bioremed. J. 9: 77–85.
Phale P.S., Sharma A., Gautam K. 2019. Microbial degradation
of xenobiotics like aromatic pollutants from the terrestrial
environments. In: Prasad M.N.V., Vithanage M., Kapley
A. Pharmaceuticals and Personal Care Products: Waste
Management and Treatment Technology. Butterworth-
Heinemann, pp. 259–278.
Pointing S. 2001. Feasibility of bioremediation by white-rot fungi.
Appl. Microbiol. Biotechnol. 57: 20–33.
Prieto A., Möder M., Rodil R., Adrian L., Marco-Urrea E. 2011.
Degradation of the antibiotics noroxacin and ciprooxacin
by a white-rot fungus and identication of degradation
A. Kathiravan, J.J. Gnanadoss
117
products. Bioresour Technol. 102: 10987–10995.
Pühse M., Szweda R.T., Ma Y., Jeworrek C., Winter R., Zorn
H. 2009. Marasmius scorodonius extracellular dimeric
peroxidase—exploring its temperature and pressure stability.
Biochim. Biophys. Acta Proteins Proteom. 1794: 1091–1098.
Radhika R., Jebapriya G.R., Gnanadoss J.J. 2013. Production of
cellulase and laccase using Pleurotus sp. under submerged and
solid-state fermentation. Int. J. Curr. Sci. 6: 7–13.
Ramachandran R., Gnanadoss J.J. 2013. Mycoremediation for the
treatment of dye containing euents. Int. J. Comput. Algor. 2:
286–293.
Ramesh C., Pattar M.G. 2009. Biodegradation of pentachlorophenol
by white-rot fungi isolated from forests of Western Ghats of
Karnataka India. Curr. Trends Biotechnol. Pharm. 3: 417–427.
Ravichandran A., Sridhar M. 2016. Versatile peroxidases: super
peroxidases with potential biotechnological applications-a
mini review. J. Dairy Vet. Anim. Res. 4: 277–280.
Risdianto H., Soanti E., Suhardi S.H., Setiadi T. 2012.
Optimization of laccase production using white rot fungi and
agricultural wastes in solid-state fermentation. J. Eng. Technol.
Sci. 44: 93–105.
Robinson T., Chandran B., Nigam P. 2001. Studies on
the production of enzymes by white-rot fungi for the
decolourisation of textile dyes. Enzyme Microb. Technol. 29:
575–579.
Rodríguez-Couto S. 2016. Potential of white-rot fungi to
treat xenobiotic-containing wastewater. In: Purchase D.
(Ed.) Fungal Applications in Sustainable Environmental
Biotechnology. Springer International Publishing Switzerland,
pp. 91–113.
Rodríguez-Couto S. 2019. Fungal laccase: a versatile enzyme
for biotechnological applications. In: Jadav A.V. et al. (Eds.)
Recent Advancement in White Biotechnology rough Fungi.
Springer Nature Switzerland, pp. 429–457.
Rubilar O., Feijoo G., Diez C., Lu-Chau T.A., Moreira M.T., Lema
J.M. 2007. Biodegradation of pentachlorophenol in soil slurry
cultures by Bjerkandera adusta and Anthracophyllum discolor.
Industr. Eng. Chem. Res. 46: 6744–6751.
Rusch M., Spielmeyer A., Zorn H., Hamscher G. 2018.
Biotransformation of ciprooxacin by Xylaria longipes:
structure elucidation and residual antibacterial activity of
metabolites. Appl. Microbiol. Biotechnol. 102: 8573–8584.
Ryu W.R., Shim S.H., Jang M.Y., Jeon Y.J., Oh K.K., Cho M.H.
2000. Biodegradation of pentachlorophenol by white-rot
fungi under ligninolytic and nonligninolytic conditions.
Biotechnol. Bioprocess Eng. 5: 211–214.
Saiu G., Tronci S., Grosso M., Cadoni E., Curreli N. 2018. Pyrene
and chrysene tolerance and biodegradation capability of
Pleurotus sajor-caju. Open Chem. Eng. J. 12: 24–35.
Salvachúa D., Prieto A., Martínez Á.T., Martínez M.J. 2013.
Characterization of a novel dye-decolorizing peroxidase
(DyP)-type enzyme from Irpex lacteus and its application in
enzymatic hydrolysis of wheat straw. Appl. Environ. Microbiol.
79: 4316–4324.
Sangale M.K., Shahnawaz M., Ade A. 2012. A review on
biodegradation of polythene: the microbial approach. J.
Bioremediat. Biodegrad. 3: 1–9.
Sen S.K., Raut S., Bandyopadhyay P., Raut S. 2016. Fungal
decolouration and degradation of azo dyes: a review. Fungal
Biol . R ev. 30: 112–133.
Serbent M.P., Guimarães D.K.S., Drechsler-Santos E.R., Helm
C.V., Giongo A., Tavares L.B.B. 2020. Growth, enzymatic
production and morphology of the white-rot fungi Lentinus
crinitus (L.) Fr. upon 2, 4-D herbicide exposition. Int. J.
Environ. Sci. Technol. 17: 2995–3012.
Serrano-González M.Y., Chandra R., Castillo-Zacarias
C., Robledo-Padilla F., Rostro-Alanis M.D.J., Parra-
Saldivar R. 2018. Biotransformation and degradation of
2,4,6-trinitrotoluene by microbial metabolism and their
interaction. Defence Technol. 14: 151–164.
Seymour R.B. 1989. Polymer science before and aer 1899:
notable developments during the lifetime of Maurits Dekker.
J. Macromol. Sci. 26: 1023–1032.
Shah V., Nerud F. 2002. Lignin degrading system of white-rot
fungi and its exploitation for dye decolorization. Can. J.
Microbiol. 48: 857–870.
Shanmugapriya S., Manivannan G., Selvakumar G., Sivakumar N.
2019. Extracellular fungal peroxidases and laccases for waste
treatment: recent improvement. In: Yadav A.N. et al. (Eds.)
Recent Advancement in White Biotechnology rough Fungi.
Springer Nature Switzerland, pp. 153–187.
Shedbalkar U., Dhanve R., Jadhav J. 2008. Biodegradation of
triphenylmethane dye cotton blue by Penicillium ochrochloron
MTCC 517. J. Hazard. Mater. 157: 472–479.
Singh B.K., Kuhad R.C. 2000. Degradation of insecticide lindane
(γ‐HCH) by white‐rot fungi Cyathus bulleri and Phanerochaete
sordida. Pest Manag. Sci. 56: 142–146.
Singh D., Gupta N. 2020. Microbial laccase: a robust enzyme and
its industrial applications. Biologia 75: 1183–1193.
Singh N., Kumar A., Sharma B. 2019. Role of fungal enzymes for
bioremediation of hazardous chemicals. In: Yadaw A.N. et al.
(Eds.) Recent Advancement in White Biotechnology rough
Fungi. Springer Nature Switzerland, pp. 237–256.
Singh S.K., Khajuria R., Kaur L. 2017. Biodegradation of
ciprooxacin by white rot fungus Pleurotus ostreatus. 3Biotech
7: 69.
Sinha S., Chattopadhyay P., Pan I., Chatterjee S., Chanda P.,
Bandyopadhyay D., Sen S.K. 2009. Microbial transformation
of xenobiotics for environmental bioremediation. Afr. J.
Biotechnol. 8: 6016–6027
Smith M.R. 1990. e biodegradation of aromatic hydrocarbons
by bacteria. Biodegradation 1: 191–206.
Sodaneath H., Lee J.I., Yang S. O., Jung H., Ryu H.W., Cho K.S.
2017. Decolorization of textile dyes in an air-li bioreactor
inoculated with Bjerkandera adusta OBR105. J. Environ. Sci.
Health A 52: 1099–1111.
Sridevi V., Lakshmi M.V.V.C., Swamy A.V.N., Rao M.N. 2011.
Implementation of response surface methodology for phenol
degradation using Pseudomonas putida (NCIM 2102). J.
Bioremediat. Biodegrad. 2: 1–7.
Suda T., Hata T., Kawai S., Okamura H., Nishida T. 2012.
Treatment of tetracycline antibiotics by laccase in the presence
of 1-hydroxybenzotriazole. Bioresour. Technol. 103: 498–501.
Sukor M.Z., Yin C.Y., Savory R.M., Abdul-Talib S. 2012.
Biodegradation kinetics of naphthalene in soil medium using
Pleurotus ostreatus in batch mode with addition of brous
biomass as a nutrient. Bioremed. J. 16: 177–184.
Tanabe J., Ishiguri F., Ohshima J., Iizuka K., Yokota S. 2016.
Changes in lignocellulolytic enzyme activity during the
degradation of Picea jezoensis wood by the white-rot fungus
Porodaedalea pini. Int Biodeterior. Biodegrad. 110: 108–112.
Teerapatsakul C., Pothiratana C., Chitradon L., achepan S.
White-rot fungi mediated bioremediation for xenobiotic degradation
118
2016. Biodegradation of polycyclic aromatic hydrocarbons
by a thermo tolerant white rot fungus Trametes polyzona
RYNF13. J. Gen. Appl. Microbiol. 62: 303–312.
Tekere M., Read J.S., Mattiasson B. 2005. Polycyclic aromatic
hydrocarbon biodegradation in extracellular uids and static
batch cultures of selected sub-tropical white-rot fungi. J.
Biotechnol. 115: 367–377.
Terrazas-Siles E., Alvarez T., Guieysse B., Mattiasson B. 2005.
Isolation and characterization of a white rot fungus
Bjerkandera sp. strain capable of oxidizing phenanthrene.
Biotechnol. Lett. 27: 84–851.
amvithayakorn P., Phosri C., Pisutpaisal N., Krajangsang S.,
Whalley A.J., Suwannasai N. 2019. Utilization of oil palm
decanter cake for valuable laccase and manganese peroxidase
enzyme production from a novel white-rot fungus,
Pseudolagarobasidium sp. PP17-33. 3Biotech 9: 417.
urston C.F. 1994. e structure and function of fungal laccases.
Microbiology 140: 19–26.
Tian Q., Dou X., Huang L., Wang L., Meng D., Zhai L., Liao X. 2020.
Characterization of a robust cold-adapted and thermostable
laccase from Pycnoporus sp. SYBC-L10 with a strong ability
for the degradation of tetracycline and oxytetracycline by
laccase-mediated oxidation. J. Hazard. Mater. 382: 121084.
Tigini V., Spina F., Romagnolo A., Prigione V.P., Varese G. 2013.
Eective biological treatment of landll leachates by means of
selected white rot fungi. Chem. Eng. Trans. 32: 265–270
Tišma M., Zelić B., Vasić-Rački Đ. 2010. White-rot fungi in
phenols, dyes and other xenobiotics treatment–a brief review.
Croatian J. Food Sci. Technol. 2: 34–47.
Torres-Farradá G., Manzano-León A.M., Rineau F., Leal M.R.,
ijs S., Jambon I., Vangronsveld J. 2019. Biodegradation
of polycyclic aromatic hydrocarbons by native Ganoderma
sp. strains: identication of metabolites and proposed
degradation pathways. Appl. Microbiol. Biotechnol. 103: 7203–
7215.
Tsujiyama S., Muraoka T., Takada N. 2013. Biodegradation of
2,4-dichlorophenol by shiitake mushroom (Lentinula edodes)
using vanillin as an activator. Biotechnol. Lett. 35: 1079–1083.
Valentin L., Lu-Chau T.A., López C., Feijoo G., Moreira M.T.,
Lema J.M. 2007. Biodegradation of dibenzothiophene,
uoranthene, pyrene and chrysene in a soil slurry reactor by
the white-rot fungus Bjerkandera sp. BOS55. Process Biochem.
42: 641–648.
Varela A., Martins C., Pereira C.S. 2017. A three-act play:
pentachlorophenol threats to the cork oak forest soils
mycobiome. Curr. Opin. Microbiol. 37: 42–149.
Varsha Y.M., Naga Deepthi C.H., Chenna S. 2011. An emphasis
on xenobiotic degradation in environmental cleanup. J.
Bioremediat. Biodegrad. S11: 001.
Vyas B.R.M., Bakowski S., Šašek V., Matucha M. 1994. Degradation
of anthracene by selected white-rot fungi. FEMS Microbiol.
Ecol. 14: 65–70.
Walter M., Boyd-Wilson K., Boul L., Ford C., McFadden D.,
Chong B., Pinfold J. 2005. Field-scale bioremediation of
pentachlorophenol by Trametes versicolor. Int. Biodeterior.
Biodegrad. 56: 51–57.
Wang S.S., Ning Y.J., Wang S.N., Zhang J., Zhang G.Q., Chen
Q.J. 2017. Purication, characterization, and cloning of an
extracellular laccase with potent dye decolorizing ability
from white rot fungus Cerrena unicolor GSM-01. Int. J. Biol.
Macromol. 95: 92–927.
Wang X., Song L., Li Z., Ni Z., Bao J., Zhang H. 2020. e
remediation of chlorpyrifos-contaminated soil by immobilized
white-rot fungi. J. Serb. Chem. Soc. 85: 857–868.
Wen X., Jia Y., Li J. 2009. Degradation of tetracycline and
oxytetracycline by crude lignin peroxidase prepared
from Phanerochaete chrysosporium – a white rot fungus.
Chemosphere 75: 1003–1007.
Wesenberg D., Kyriakides I., Agathos S.N. 2003. White-rot fungi
and their enzymes for the treatment of industrial dye euents.
Biotechnol. Adv. 22: 161–187.
Wirasnita R., Hadibarata T. 2016. Potential of the white-rot
fungus Pleurotus pulmonarius F043 for degradation and
transformation of uoranthene. Pedosphere 26: 49–54.
Woiciechowski A.L., de Souza Vandenberghe L.P., Karp S.G., Letti
L.A.J., de Carvalho J.C., Medeiros A.B.P., Soccol C.R. 2013.
e pretreatment step in lignocellulosic biomass conversion:
current systems and new biological systems. In: Faraco V.
(Ed.) Lignocellulose Conversion: Enzymatic and Microbial
Tools for Bioethanol Production. Springer, Berlin, pp. 3–64.
Wolfand J.M., LeFevre G.H., Luthy R.G. 2016. Metabolization
and degradation kinetics of the urban-use pesticide pronil
by white rot fungus Trametes versicolor. Env iron. Sci. Proc.
Impacts 18: 1256–1265.
Wu J., Xiao Y.Z., Yu H.Q. 2005. Degradation of lignin in pulp mill
wastewaters by white-rot fungi on biolm. Bioresour Technol.
96: 1357–1363.
Xiao P., Mori T., Kamei I., Kondo R. 2011. A novel metabolic
pathway for biodegradation of DDT by the white-rot fungi,
Phlebia lindtneri and Phlebia brevispora. Biodegradation 22:
859–867.
Xiao P., Mori T., Kamei, I., Kiyota H., Takagi K., Kondo, R. 2011.
Novel metabolic pathways of organochlorine pesticides
dieldrin and aldrin by the white-rot fungi of the genus Phlebia.
Chemosphere 85: 218–224.
Xiao P., Kondo R. 2019. Biodegradation and bioconversion of
endrin by white-rot fungi, Phlebia acanthocystis and Phlebia
brevispora. Mycoscience 60: 255–261.
Xiao P., Kondo R. 2020. Biodegradation and biotransformation
of pentachlorophenol by wood-decaying white rot fungus
Phlebia acanthocystis TMIC34875. J. Wood Sci. 66: 1–11.
Xu F. 1999. Recent progress in laccase study: properties,
enzymology, production, and applications. In: Flickinger
M.C., Drew S.W. (Eds.) Encyclopedia of Bioprocessing
Technology: Fermentation, Biocatalysis, and Bioseparation.
John Wiley & Sons, New York.
Yadav J.S., Reddy C.A. 1993. Degradation of benzene, toluene,
ethylbenzene, and xylenes (BTEX) by the lignin-degrading
basidiomycete Phanerochaete chrysosporium. Appl. Environ.
Microbiol. 59: 756–762.
Yadav M., Yadav H.S. 2015. Applications of ligninolytic enzymes
to pollutants, wastewater, dyes, soil, coal, paper and polymers.
Environ. Chem. Lett. 13: 309–318.
Yang J., Lin Y., Yang X., Ng T.B., Ye X., Lin J. 2017. Degradation
of tetracycline by immobilized laccase and the proposed
transformation pathway. J. Hazard. Mater. 322: 525-531.
Zhang H., Zhang S., He F., Qin X., Zhang X., Yang Y. 2016.
Characterization of a manganese peroxidase from white-rot
fungus Trametes sp. 48424 with strong ability of degrading
dierent types of dyes and polycyclic aromatic hydrocarbons.
J. Hazard. Mater. 320: 265–277.
Zhang X., Yan K., Ren D., Wang H. 2007. Studies on quinoline
A. Kathiravan, J.J. Gnanadoss
119
biodegradation by a white rot fungus (Pleurotus ostreatus BP)
in liquid and solid state substrates. Fresenius Environ. Bull. 16:
632–638.
Zhang Y., G eißen S.U. 2010. In vitro degradation of carbamazepine
and diclofenac by crude lignin peroxidase. J. Hazard. Mater
176: 1089–1092.
Zhang H., Zhang J., Zhang X., Geng A. 2018. Purication and
characterization of a novel manganese peroxidase from
white-rot fungus Cerrena unicolor BBP6 and its application
in dye decolorization and denim bleaching. Process. Biochem.
66: 222–229.
Zheng F., An Q., Meng G., Wu X.J., Dai Y.C., Si J., Cui B.K. 2017.
A novel laccase from white rot fungus Trametes orientalis:
Purication, characterization, and application. Int. J. Biol.
Macromol. 102: 758–770.
White-rot fungi mediated bioremediation for xenobiotic degradation
Received 28 May 2021; received in revised form 26 June 2021; accepted 10 September 2021
... Попадая в организм человека различными путями, в основном через потребление пищи, воды, вдыхание частиц микропластика, фталаты могут оказывать негативное влияние на здоровье человека разрушая гормоны тем самым негативно влияя на всю эндокринную систему человека [3,4]. При этом базидиомицеты белой гнили известны как эффективные биодеструкторы различных ксенобиотиков [5,6]. Учитывая вышеизложенное, изучение способности базидиомицетов белой гнили разлагать фталаты, а также выявление механизмов деструкции фталатов, является крайне актуальной задачей. ...
... Among all filamentous fungi, the fungi that cause white rot of wood (i.e., white-rot fungi) are of particular interest due to their unique ability to effectively metabolize many types of highly recalcitrant phenolic compounds, including aromatic dyes [12][13][14]. The first stages of the metabolic assimilation of aromatic compounds by white-rot fungi are their extracellular and intracellular oxidative degradation. ...
Article
Full-text available
As a toxic xenobiotic compound, the anthraquinone dye Remazol Brilliant Blue R (RBBR) poses a serious threat to aquatic ecosystems. In the present study, the ability of Trametes hirsuta to remove RBBR from the medium was investigated, and the role of adsorption by fungal mycelium and biodegradation by fungal enzymes was evaluated. It was shown that the whole fungal culture was able to remove up to 97% of the dye within the first four hours of incubation. Based on enzymatic activities in the culture broth, laccases were proposed to be the main enzymes contributing to RBBR degradation, and RT-qPCR measurements demonstrated an increase in transcription for the two laccase genes—lacA and lacB. Composite mycelial pellets of T. hirsuta with improved adsorption ability were prepared by adding activated carbon to the growth medium, and the induction of laccase activity by carbon was shown. For composite pellets, the RBBR decolorization degree was about 1.9 times higher at 1 h of incubation compared to carbon-free pellets. Hence, it was shown that using fungal mycelium pellets containing activated carbon can be an effective and economical method of dye removal.
... Fungi use many different enzymes to metabolize xenobiotics from the environment [36]. The saprotrophic fungus Penicillium griseofulvum is highly tolerant to xenobiotics like hexachlorocyclohexane (HCH) and toxic metals like vanadium [37]. ...
... In biobeds, the presence of lignocellulosic materials reduces the pH in the system generating an environment that favors the growth and development of lignin-degrading fungi, such as different species of white-rot fungi [45]. White-rot fungi are organisms broadly reported in the biodegradation of several organic pollutants, including pesticides from different chemical families [93][94][95][96][97][98]. Fungi can produce extracellular enzymes, such as peroxidases, laccases, and the cytochrome P450 complex, implicated in pesticide degradation [99][100][101]. ...
Article
Full-text available
Pesticides are chemical molecules employed to protect crops from pests in agriculture. The use of pesticides significantly enhances crop yields and helps to guarantee the quality of farm products; due to this, each year, millions of tons of pesticides are employed in crop fields worldwide. However, the extensive use of pesticides has been related to environmental pollution, mainly in soils and water bodies. The presence of pesticides in the environment constitutes a menace to biodiversity, soil fertility, food supply, and human health. Activities related to pesticide use in crops, such as the handling and pesticide dissolution before application, the filling and cleaning of aspersion equipment and machinery, accidental spills in crop fields, and the inadequate disposal of pesticide residues have been identified as important punctual pesticide pollution sources. Therefore, avoiding releasing pesticide residues into the soil and water is crucial to mitigating the environmental pollution associated with agricultural practices. Biobeds are biological systems that have been proposed as feasible, low-cost, and efficient alternatives for punctual pesticide pollution mitigation. Biobeds were first described as trenches packed with a mixture of 50% wheat straw, 25% soil, and 25% peat, covered with a grass layer; this composition is known as a “biomixture”. In biobeds, the biomixture absorbs the pesticide residues and supports the development of different microorganisms, such as bacteria and fungi, needed for pesticide degradation in the system. The effectiveness of a biobed systems lies in the high pesticide retention in the biomixture and the degradation potential of the microorganisms growing in the system. In this review, 24 studies published in the last five years (2018–2022) related to pesticide biodegradation in biobed systems are analyzed, emphasizing alternative biomixture composition usage, microbiological strategies, and the key physicochemical parameters for efficient pesticide degradation in the biobed systems. The availability of robust scientific evidence about the simple applicability, low cost, and effectiveness of biobeds for pesticide residue treatment is crucial to increasing the use of biobeds by farmers in different agricultural regions around the world.
... White-rot basidiomycetes degrade a wide range of xenobiotics through the action of a nonspecific ligninmodifying enzyme (LME) system that includes extracellular enzymes, such as laccase, lignin peroxidase, manganese peroxidase, and H 2 O 2 -generating enzymes [28,29]. Degradation of various PAH compounds, antibiotics, pesticides, and industrial dyes by various white-rot fungi has been well studied. ...
... MnP enzyme oxidize PAHs indirectly followed by lignin peroxide mediation of Bjerkandera sp. degraded anthracene, pyrene, and dibenzothiophene (Kathiravan and Gnanadoss 2021). Laccase and manganese dependent peroxidase produced by Peniophora incarnata degrades 86.5% of phenanthrene, 77.4% of fluoranthene, and 82.6% of pyrene and MnP of Irpexlacteus fungus degrades oligocyclic aromatic hydrocarbon more efficiently (Zainith et al. 2020). ...
Article
Full-text available
Manganese peroxidase (MnP) plays an important role in the treatment of environmental pollutants existing around us. This review article covers all its recent potential applicability in a wide range of different areas, such as textile and food industry, in alcohol production, pulp and paper industry, in biofuel, in agriculture, in cosmetic wastes and in waste water treatment. MnP plays a potential role in biodegradation of phenolic compounds and non-phenolic substrates, dyes, and many xenobiotic compounds and in bioethanol production. This
Article
A class of organic priority pollutants known as PAHs is of critical public health and environmental concern due to its carcinogenic properties as well as its genotoxic, mutagenic, and cytotoxic properties. Research to eliminate PAHs from the environment has increased significantly due to awareness about their negative effects on the environment and human health. Various environmental factors, including nutrients, microorganisms present and their abundance, and the nature and chemical properties of the PAH affect the biodegradation of PAHs. A large spectrum of bacteria, fungi, and algae have ability to degrade PAHs with the biodegradation capacity of bacteria and fungi receiving the most attention. A considerable amount of research has been conducted in the last few decades on analyzing microbial communities for their genomic organization, enzymatic and biochemical properties capable of degrading PAH. While it is true that PAH degrading microorganisms offer potential for recovering damaged ecosystems in a cost-efficient way, new advances are needed to make these microbes more robust and successful at eliminating toxic chemicals. By optimizing some factors like adsorption, bioavailability and mass transfer of PAHs, microorganisms in their natural habitat could be greatly improved to biodegrade PAHs. This review aims to comprehensively discuss the latest findings and address the current wealth of knowledge in the microbial bioremediation of PAHs. Additionally, recent breakthroughs in PAH degradation are discussed in order to facilitate a broader understanding of the bioremediation of PAHs in the environment.
Article
Full-text available
Alkene cleavage is a possibility to generate aldehydes with olfactory properties for the fragrance and flavor industry. A dye-decolorizing peroxidase (DyP) of the basidiomycete Pleurotus sapidus (PsaPOX) cleaved the aryl alkene trans-anethole. The PsaPOX was semi-purified from the mycelium via FPLC, and the corresponding gene was identified. The amino acid sequence as well as the predicted tertiary structure showed typical characteristics of DyPs as well as a non-canonical Mn2+-oxidation site on its surface. The gene was expressed in Komagataella pfaffii GS115 yielding activities up to 142 U/L using 2,2'-azino-bis(3-ethylbenzthiazoline-6-sulphonic acid) as substrate. PsaPOX exhibited optima at pH 3.5 and 40 °C and showed highest peroxidase activity in the presence of 100 µM H2O2 and 25 mM Mn2+. PsaPOX lacked the typical activity of DyPs towards anthraquinone dyes, but oxidized Mn2+ to Mn3+. In addition, bleaching of β-carotene and annatto was observed. Biotransformation experiments verified the alkene cleavage activity towards the aryl alkenes (E)-methyl isoeugenol, α-methylstyrene, and trans-anethole, which was increased almost twofold in the presence of Mn2+. The resultant aldehydes are olfactants used in the fragrance and flavor industry. PsaPOX is the first described DyP with alkene cleavage activity towards aryl alkenes and showed potential as biocatalyst for flavor production.
Article
Full-text available
Dichlorophenoxyacetic acid (2,4-D) is one of the most commonly used weed control herbicides. White-rot fungi (WRF) are recognized as a degrader of a wide range of molecules because due to their enzymatic plasticity, nevertheless, the knowledge of their ability to degrade 2,4 D residues in the environment is still limited. In this study, the tolerance and the mycelium growth kinetics of twelve WRF to 2,4-D (670 g L−1) were evaluated on potato dextrose agar added with 2,4-D using the technique of cup plate. Agaricus sp., Pleurotus pulmonarius, Pleurotus djamor and the EF 58 strain grew in all plates in the presence of 2,4-D. Molecular phylogenetic analysis of the EF 58 strain allowed its determination as a lineage belonging to Lentinus crinitus Fr. (Polyporales, Basidiomycota) species-complex. The daily fungal growth rate of L. crinitus EF 58 lineage in medium with (5.025 g L−1 and 50.25 g L−1) and without the addition of the herbicide, laccase production and mycelium structure after 6, 10 and 15 days of exposure to 2,4-D was analyzed. There were no significant differences between the control and C1 treatments concerning the growth of L. crinitus EF 58 lineage. The enzymatic activity tests showed evidence for the presence of laccases in all essays. Optical microscopy observations did not reveal substantial alterations in its mycelium morphology after exposition to the herbicide. It was the first study showing the potential of L. crinitus EF58 lineage in tolerating 2,4-D what represents the beginning of work for the bioremediation process with practical applications.
Article
Full-text available
Pentachlorophenol (PCP) has been introduced into the environment mainly as a wood preservative and biocide. The degradation and transformation of PCP in liquid culture by wood-decaying fungus capable of degrading organochlorine pesticides was investigated in this study. The results of tolerance test showed that the tolerance level of Phlebia acanthocystis to PCP in potato dextrose agar medium was higher than that of other Phlebia species. At the end of 10 days of incubation, P. acanthocystis was able to remove 100% and 76% of PCP (25 μM) in low-nitrogen and potato dextrose broth media, respectively. The decrease of PCP in P. acanthocystis culture is accompanied by the formation of pentachloroanisole and p -tetrachlorohydroquinone via methylation and oxidation reactions. Moreover, the p -tetrachlorohydroquinone formed is rapidly converted to methylated products including tetrachloro-4-methoxyphenol and tetrachloro-1,4-dimethoxybenzene. The activities of lignin peroxidase and manganese peroxidase were found to increase in extracellular fluid from fungal culture treated with high-concentration PCP, with maximum values of 169.6 U/L and 73.4 U/L, respectively. The in vitro degradation of PCP and p -tetrachlorohydroquinone was confirmed using extracellular fluid of P. acanthocystis , suggested that the methylation of both compounds is related to extracellular enzymes. Degradation of PCP was efficiently inhibited by piperonyl butoxide or 1-aminobenzotriazole, demonstrating that cytochrome P450 monooxygenase is involved in fungal transformation of PCP, particularly in the oxidation of PCP to p -tetrachlorohydroquinone. Additionally, P. acanthocystis mineralized 9.3% of the PCP to ¹⁴ CO 2 in low-nitrogen culture during 42 days. Results obtained in the present study are in favor of the use of P. acanthocystis as a microbial tool of remediation of PCP-contaminated sites.
Article
Full-text available
Laccases (E.C. 1.10.3.2) are most stable and powerful biocatalysts with broad applications. It is a copper containing, organic solvents resistance, oxidoreductase enzyme, having potential ability to oxidize the phenolic and non-phenolic compounds to produce dimers, oligomers, and polymers. Laccase are widely distributed in nature ranging from prokaryotes to lower eukaryotes and fungi to plants. In their natural surroundings, Laccases are responsible for the biological degradation of lignin and received much attention among researchers in last decade, due to their wide-spread applicability and stability. The huge versatility of laccase due to their broad range of substrate specificity makes laccase highly suitable for various fields of industrial applications such as paper and pulp industry, textile applications, bioremediation, food processing, pharmaceutical and nanobiotechnological applications, bio-bleaching of synthetic dyes and biodegradation of xenobiotic compounds. Laccase mediated enzymatic treatment has been considered as, an effective process towards their adequate effect of real industrial decolourization of dye-contaminated wastewater, improving and texturing the food quality in affordable cost, less-time consuming at state-of-environmental risk. A present review is a brief analysis about this prominent group of enzymes and increase knowledge about its applications further which would encourage the use of laccase based production for economical purposes in future.
Article
Full-text available
This research focused on the degradation of chlorpyrifos via immobilized white rot fungi in soil, with the aim to select excellent degrading strains and an optimal carrier of white rot fungi. Immobilization of white rot fungi was assessed on corn stover, wheat straw, peanut shells, wood chip, and corn cobs. Phlebia sp., Lenzites betulinus and Irpex lacteus were grown in defined nutrient media for the remediation of pesticide-contaminated soils. The carrier of the biomass was determined by observing the growth of white rot fungi. The results showed that corn stover and wheat straw are suitable carriers of the immobilized white rot fungi and that Phlebia sp. and Lenzites betulinus have a positive effect on the degradation of chlorpyrifos. At 30°C and neutral pH, the degradation rate of chlorpyrifos was 74.35 %, Phlebia sp. being immobilized by corn stover in 7 days, which was the best result compared to other combinations of strains and carriers. The orthogonal experiment showed that the pH value and temperature affected the pollutant degradability more than the initial concentration and the biomass dosage.
Article
The laccase enzyme from white-rot fungus Trametes hirsuta brought about degradation of 0.5 mg/L chloramphenicol within seven days without mediators, which is typically found in the pharmaceutical wastes. This study is first of its kind which showed that the higher concentrations of chloramphenicol (10 mg/L) was efficiently degraded in the presence of laccase mediator system (LMS) like syringaldehyde, vanillin, ABTS and α-naphthol. The addition of mediators escalated the degradation process from 10% to 100% within 48 h. The kinetic parameter Vmax of the laccase mediated degradation reactions with and without mediators comprehensively proved the potentiality of laccase and its redox mediators in bioremediation of most stable micro-pollutants i.e., antibiotics in the pharmaceutical wastes. Further, the laccase mediated degradation of chloramphenicol was confirmed by the microbial susceptibility assay. The formation of chloramphenicol aldehyde after degradation of chloramphenicol was confirmed through LC-MS analysis and proved to be non-toxic to the tested micro-organisms.
Article
Oil palm decanter cake (OPDC) in the current study was converted to valuable products as laccase and manganese peroxidase (MnP) by an undescribed strain of the white-rot fungus, Pseudolagarobasidium sp. PP17-33. The optimization to enhance the production of enzymes through solid-state fermentation was performed using Plackett–Burman design and response surface methodology. The highest observed laccase was 5.841 U/gds and observed MnP was 5.156 U/gds, which enhanced yield by 2.59-fold and 1.94-fold from the non-optimization. The optimized medium (mg/g of OPDC) consisted of 0.852 mg CuSO4·5H2O, 13.512 mg glucose, 2 mg yeast extract, 0.2 mg KH2PO4, 1.5 mg MgSO4·7H2O, 0.01 mg FeSO4·7H2O, 0.15 mg MnSO4·H2O, 0.01 mg ZnSO4·7H2O and 0.3 mg Tween 80 (pH 5.0) when incubated at 30 °C for 7 days. The most significant variables of laccase and MnP productions were CuSO4·5H2O and glucose concentrations. This study is the first to report on the production of ligninolytic enzymes from OPDC waste using white-rot fungi. In addition, five different white-rot fungi, Coriolopsis aspera, C. retropicta, Dentipellis parmastoi, Nigroporus vinosus and Tyromyces xuchilensis, are newly observed producers of ligninolytic enzymes in Thailand. The results obtained from this study are significant not only for agro-industrial waste management but also for value-added enzyme production.
Chapter
Environmental hazard is growing more and more due to the indiscriminate and frequently deliberate release of harmful substances. Use of chemicals in industrial processes including nuclear experiments, agricultural practices, and various aspects of our daily lives resulted into the release of potential hazardous chemicals into the environment either on purpose or by accident. These hazardous chemicals known to pollute the environment are pesticides, heavy metals, hydrocarbons, drugs, halogenated solvents, and agricultural chemicals. After their release into environment, these chemicals are transported through the water, soil, and atmosphere sources. Fungi play a very crucial role in bioremediation of hazardous chemicals owing to their robust morphology and diverse metabolic capacity. Fungal enzymes have potential to effectively transform and detoxify hazardous substances. They have been recognized to be able to transform pollutants at a detectable rate and are potentially suitable to restore polluted environments. The fungal degradation of xenobiotics is looked upon as an effective method of removing these pollutants from the environment by a process which is currently known as bioremediation. The present chapter focuses on different fungal groups secreted a number of enzymes from a variety of habitats with their role in bioremediation of different toxic and recalcitrant compounds. This chapter presents an extensive review of the fungal activities on hazardous chemicals, fungal diversity, and the use of fungi in the degradation of chemical pollutants, enzyme degrading systems, and perspectives on the use of fungi in bioremediation and unexplored research.
Chapter
The major thrust of scientific research is pollution control due to increased discharge and improper management of industrial wastes, especially textile industries. Textile industry is one of the major industries, which uses many xenobiotics as dyes and releases several undesirable pollutants into the environment. A wide variety of dyes were used in the textile industry, which are complex structured and constitute the largest group among the recalcitrant xenobiotics. Due to lower degree of dye fixation to fabrics, more than 10% of the dyes goes into wastewater and released into the environment unaltered. Dye removal can be done with physical and physicochemical methods, but these methods are expensive and require operation expertise. Complete breakdown of the dye molecules is the desired outcome and that is possible with biological means. Decolourization with biological means has gained great attention, and many researchers suggested several biotechnological approaches for combating the textile pollution. Many bacteria are having enzymes for complete degradation of the azodyes, but it needs alterations in the process. Recently, fungal decolourization, especially white rot fungi, is gaining importance, and these fungi are capable of producing one or more extracellular, non-specific, non-selective enzymes which can able to degrade a wide range of xenobiotics. The white rot fungal enzymes are mainly composed of lignin peroxidase, manganese-dependent peroxidases, laccases and hydrogen peroxide-producing peroxidases. They are the most efficient microorganisms degrading textile dyes, which are structurally different and complex. White rot fungal enzymes and its degradation abilities to remove synthetic dyes from textile wastewater are compiled in this chapter.